You are on page 1of 17

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/216810670

Measuring Net Primary Production in Forests: Concepts and Field Methods

Article  in  Ecological Applications · April 2001


DOI: 10.2307/3060894

CITATIONS READS

699 5,673

6 authors, including:

Deborah A Clark Sandra Brown


University of Missouri - St. Louis Winrock International
124 PUBLICATIONS   11,879 CITATIONS    230 PUBLICATIONS   31,549 CITATIONS   

SEE PROFILE SEE PROFILE

David W. Kicklighter John Richard Thomlinson


Marine Biological Laboratory California State University, Dominguez Hills
194 PUBLICATIONS   18,185 CITATIONS    32 PUBLICATIONS   2,757 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

A meta-analysis about impacts of climatic change on vegetation in China View project

Carbon transportation and transformation among hydrosphere, lithosphere, pedosphere, atmosphere and biosphere in karst region: model coupling and regulated
management View project

All content following this page was uploaded by Deborah A Clark on 17 May 2014.

The user has requested enhancement of the downloaded file.


Measuring Net Primary Production in Forests: Concepts and Field Methods

Deborah A. Clark; Sandra Brown; David W. Kicklighter; Jeffrey Q. Chambers; John R.


Thomlinson; Jian Ni

Ecological Applications, Vol. 11, No. 2. (Apr., 2001), pp. 356-370.

Stable URL:
http://links.jstor.org/sici?sici=1051-0761%28200104%2911%3A2%3C356%3AMNPPIF%3E2.0.CO%3B2-Q

Ecological Applications is currently published by Ecological Society of America.

Your use of the JSTOR archive indicates your acceptance of JSTOR's Terms and Conditions of Use, available at
http://www.jstor.org/about/terms.html. JSTOR's Terms and Conditions of Use provides, in part, that unless you have obtained
prior permission, you may not download an entire issue of a journal or multiple copies of articles, and you may use content in
the JSTOR archive only for your personal, non-commercial use.

Please contact the publisher regarding any further use of this work. Publisher contact information may be obtained at
http://www.jstor.org/journals/esa.html.

Each copy of any part of a JSTOR transmission must contain the same copyright notice that appears on the screen or printed
page of such transmission.

JSTOR is an independent not-for-profit organization dedicated to and preserving a digital archive of scholarly journals. For
more information regarding JSTOR, please contact support@jstor.org.

http://www.jstor.org
Mon Jun 4 16:32:00 2007
Ecoiog~calA pplicarions, 1 1(2), 2001, pp. 356-370
C 2001 by the Ecological Society of America

MEASURING NET PRIMARY PRODUCTION IN FORESTS:

CONCEPTS AND FIELD METHODS

A . CLARK,'SANDRABROWN,',' DAVIDW . KICK LIGHTER,^ JEFFREY


DEBORAH Q . CHAMBERS,"
J O H N R. T H O M L I N S OA N
N D, ~J I A N N16

'Department of Biology, University of Missouri-St. Louis; Mailing address: INTERLINK-341, P. 0 . Box 02-5635,

Miami, Florida 33102 USA

2Department of Natural Resources and Environmental Sciences, University of Illinois, W503 Turner Hall,

Urbana, Illinois 61801 USA

'The Ecosystems Center, Marine Biological Laboraton, Woods Hole, Massachusetts 02543 USA

'National Center for Ecological Analysis and Synthesis, University of California,

Sanra Barbara, California 93101-3351 USA

51nstitutefor Tropical Ecosystem Studies, University of Puerto Rico, P.O. Box 363682,

Sun Juan, Puerto Rico 00936-3682 USA

hLaboraton of Quantitative Vegetation Ecology, Institute of Botany, Chinese Academy of Sciences,

Xiangshan Nanxincun 20, 100093 Beijing, P.R. China

Abstract. There are pressing reasons for developing a better understanding o f net
primary production (NPP) in the world's forests. These ecosystems play a large role in the
world's carbon budget, and their dynamics, which are likely to be responding to global
changes in climate and atmospheric composition, have major economic implications and
impacts on global biodiversity. Although there is a long history o f forest NPP studies in
the ecological literature, current understanding o f ecosystem-level production remains lim-
ited. Forest NPP cannot be directly measured; it must be approached by indirect methods.
To date, field measurements have been largely restricted to a few aspects o f NPP; methods
are still lacking for field assessment o f others, and past studies have involved confusion
about the types o f measurements needed. As a result, existing field-based estimates o f forest
NPP are likely to be significant underestimates.
In this paper we provide a conceptual framework to guide efforts toward improved
estimates o f forest NPP. W e define the quantity NPP* as the summed classes o f organic
material that should be measured or estimated in field studies for an estimate o f total NPP.
W e discuss the above- and belowground components o f NPP* and the available methods
for measuring them in the field. W e then assess the implications o f the limitations o f past
studies for current understanding o f NPP in forest ecosystems, discuss how field NPP*
measurements can be used to complement tower-based studies o f forest carbon flux, and
recommend design criteria for future field studies o f forest NPP.
Key words: biomass increment; boreal, temperate, and tropical forests; carbon; coarse roots;
fine root turnover; forest inventon plots; litterfall; net ecosystem C exchange; net p r i m a n production;
total belowground carbon allocation.

concentrations. At the same time, changing climate and


An important current research need is to develop a atmospheric chemistry (e.g., CO, levels, nitrogen de-
better understanding o f net primary production (NPP) position) are likely to be causing on-going changes in
in the world's forests, ecosystems that play a major forest NPP. The design and evaluation o f global-scale
role in the global carbon budget (Dixon et al. 1994). carbon models require field estimates o f forest NPP
While unprecedented atmospheric concentrations (Petit and how it is responding to these global changes. In
et al. 1999) o f the greenhouse gas carbon dioxide (CO,) addition, a better grasp o f NPP would help improve
continue to increase due to anthropogenic activities, assessments o f forest-level carbon ( C ) exchange with
large uncertainties affect current understanding o f the the atmosphere developed from eddy covariance mea-
world's carbon budget (Melillo et al. 1996). One such surements ( c f . Goulden et al. 1998, Lindroth et al.
uncertainty is the balance between NPP and heterotro- 1998, Running et al. 1999). Such improvements in our
phic respiration in forests globally. Small shifts be- understanding o f forest carbon dynamics can then be
tween these fluxes can greatly affect atmospheric CO, used to develop better policy decisions related to forest
production or conservation.
Manuscript received 29 November 1999; revised 16 March Progress in understanding NPP and its controls in
2000; accepted 10 April 2000: final version received 10 April
2000.
forest ecosystems is hindered by the limitations o f the
' Present address: Winrock International, 161 1 N. Kent existing field data (see recent reviews for boreal and
Street, Suite 600, Arlington, Virginia 22209 USA. tropical forests, respectively: Clark et al. 2001, Gower
April 2001 MEASURING FOREST NPP

a NPP b NPP* [THE MEASUREMENT]


(ALL ORGANIC MATERIAL FIXED
IN THE INTERVAL)

INCREMENTS + LOSSES
[Net increments in above-and [Materialslostlshed during the interval,
belowground live biomass] that represent additional production]
ABOVEGROUND I I

ABOVEGROUND
NEW ABOVEGROUND BIOMASS
(newwood [stemslbranches],new leaves.
new stores of non-structural CHOW
/ NEW REPRODUCTIVE MATERIALS
(inflorescences,fruitslseeds,nectar)
INCREMENT
(net increases in wood in stems &
FlNE LITTERFALL
NEW VOLATILE & LEACHABLE ORGANICS (shed leaves,twigs, fruitslflowers)
LOSSES TO CONSUMERS
( herbivory,frugivory,sap-sucking)
BELOWGROUND

NEW COARSE ROOT BIOMASS


I (incl.new stores of non-structuralCHO's) I
NET COARSE ROOT INCREMENT DEAD COARSE ROOTS
I NEW FlNE ROOT BIOMASS I NET FlNE ROOT INCREMENT DEAD FlNE ROOTS
NEW ROOT EXUDATES
I ROOT LOSSES TO HERBIVORES 1 I
NEW CHOS EXPORTED TO SYMBIONTS
(rootnodules, mycorrhizae) ROOT EXUDATES II I
CHO EXPORT TO SYMBIONTS
(rootnodules, mycorrhizae)

FIG. 1. The components of (a) forest NPP and (b) NPP*, the sum of all materials that together represent: ( I ) the amount
of new organic matter that is retained by live plants at the end of the interval, and (2) the amount of organic matter that was
both produced and lost by the plants during the same interval. CHO = carbohydrates.

et al. 2001). Although the ecological literature contains however, it is not possible to measure forest NPP in
a plethora of papers on the topic, reported estimates of terms of this difference (Waring and Schlesinger 1985).
forest NPP are based on incomplete, and sometimes GPP cannot be measured directly (Ryan 1991), and
inappropriate, field measurements. The substantial ef- estimating total plant respiration at the ecosystem level
forts that are required for NPP field studies, the un- remains difficult and involves significant uncertainties
resolved methods challenges, and a frequent lack of (cf. Ryan et al. 1996, Lavigne et al. 1997). Alterna-
conceptual clarity are continuing obstacles to improv- tively, NPP is defined as the total new organic matter
ing our knowledge of NPP. produced during a specified interval. Although the
In this paper we examine how forest NPP (above- components of this production are readily conceptu-
and belowground) can be estimated based on field mea- alized (Fig. la), they cannot be directly measured in
surements. We then assess the implications of the lim-
the field because of transformations (consumption, de-
itations of past studies for current perceptions of the
composition, mortality, export) they undergo during the
amount of NPP in forests globally, and we discuss how
measurement interval. Instead, NPP must be estimated
field NPP studies can be used to evaluate tower-based
measurements of forest carbon flux. We conclude with based on a suite of measurements of various types and
a set of priorities and design criteria for field studies numerous underlying assumptions. To clarify the un-
aimed at a more complete assessment of forest NPP derlying concepts and to provide a complete and in-
than has been achieved to date. ternally consistent framework for field studies, we de-
fine the quantity NPP*, the field-measurement-based,
ESTIMATING
FORESTNPP I N THE FIELD operational estimate of actual NPP. NPP* (Fig. Ib) is
Net primary production is the difference between the sum of all materials that together are equivalent to:
total photosynthesis (Gross Primary Production, GPP) ( I ) the amount of new organic matter that is retained
and total plant respiration in an ecosystem. In the field, by live plants at the end of the interval, and (2) the
Ecological Applications
DEBORAH A. CLARK ET AL.
Vol. 1 1 , No. 2

amount of organic matter that was both produced and for estimates of aboveground increment. In smaller
lost by the plants during the same interval. stature stands such as young second-growth or tropical
In practice, few NPP* components are measured in dry forest, a lower minimum tree size should be used
field studies in forest ecosystems. Most frequently, and documented (cf. Murphy and Lugo 1986, Brown
measurements are restricted to fine litterfall and above- 1997).
ground biomass increment, and their sum is considered Aboveground biomass increment is estimated from
equivalent to aboveground NPP (ANPP). Belowground two successive stand-level biomass estimates. Biomass
components are often ignored or are estimated as some is estimated by applying harvest-based allometric re-
theoretical proportion of aboveground values. For ex- gression equations to measurements of the diameters
ample, among 48 field NPP studies in tropical forests of all trees in a plot that are above the minimum size.
(Clark et al. 2001: Appendix), litterfall and above- Developing site-specific allometric equations for forest
ground biomass increment were measured in 94% and trees is laborious (harvesting one tropical emergent can
6096, respectively, of the studies; no other component require >25 person-days). Researchers therefore com-
of aboveground NPP* (Fig. lb) was measured in any monly use existing allometric equations (cf. Brown
study, and in only 13% of the studies was any aspect 1997, Gower et al. 1999). Because of the potential for
of belowground production assessed. Similarly, Long intersite variation in factors such as tree architecture
and Hutchin (1991) reported that in < l o % of the In- and wood density, this practice can introduce errors in
ternational Biological Programme NPP studies was any estimated aboveground increment (see Gower et al.
belowground biomass measured. In addition, data re- 1999). For example, Grier et al. (1984) used both site-
ported in some past NPP studies are unusable due to specific and generalized regression equations to cal-
inadequate methods and/or inadequate documentation culate foliage biomass for five Pseudotsuga rnenziesii
of methods. stands, and found the generalized equations produced
Reliable assessment of the amount of forest NPP will errors of -24 to +93%. For this reason, when locally
require quantifying all materials that contribute to total derived equations for the species under study are not
NPP* (Fig. lb), for at least a benchmark set of sites. available, it is important to match the allometric equa-
When complete accounting is available for represen- tion as closely as possible to the site under study (i.e.,
tative sites in each major forest type, it will be possible use an equation based on data from one or more sites
to identify those materials that can be ignored without of comparable climatic and edaphic conditions), or,
producing serious underestimates of site NPP. For all preferably, to test its predictions by first measuring and
NPP* components in forests, however, there are prac- then harvesting individual trees or forest plots on site.
tical and theoretical challenges for obtaining accurate Two approaches (Fig. 2) can then be used for esti-
estimates. Further, the appropriate methods for some mating aboveground biomass increment from field
of them will differ among forest types. Below we re- measurements and biomass allometry. Both approaches
view these considerations for all NPP* constituents. give the same estimated aboveground biomass incre-
ment, although tree mortality and ingrowth (trees that
Aboveground increments a n d losses grew past the minimum diameter during the interval)
Aboveground biornass increment.-In most forest have to be accounted for differently in the two cases.
ecosystems, aboveground biomass and its increment Not understanding the subtleties between these two
are strongly dominated by the overstory trees. For ex- methods can lead to erroneous estimates of above-
ample, it has been estimated that understory vegetation ground biomass increment and thus ANPP or total NPP.
in mature moist tropical forests generally comprises Approach 1 is based on tracking individual trees.
<3% of the aboveground biomass (Brown 1997); given The increment for each tree is calculated as the dif-
the low light levels close to the ground in these forests, ference between its estimated biomass at the beginning
the contribution by this stratum to total aboveground and end of the interval. If a tree dies in the measurement
biomass increment is negligible. Important exceptions interval and the intercensus interval is short, the tree
to this general rule are boreal forests and temperate and can be assumed to have no increment and is ignored
tropical woodlands with an open overstory and a dense in the calculation (but see Approach 2). Thus, for the
ground cover or shrub layer. In such forests, where stand, increments are summed for all trees surviving
production in the understory can be substantial and may the interval. This total is then adjusted for ingrowth;
even exceed that of the trees (cf. Black et al. 1996, the increment of each new tree is calculated as the
Gower et al. 2001), estimating total NPP requires spe- difference between its estimated biomass at the end of
cial techniques for measurement of aboveground pro- the interval and the biomass of a tree of the minimum
duction by nontree vegetation. In most closed forests, measured diameter. The summed increments of the in-
however, aboveground biomass production can be re- growth are then added to the stand increment. A variant
liably based on the biomass increment by trees above of this approach, but with the same calculation methods
a carefully chosen minimum size. In large-stature for- (Approach 1, Fig. 2), can be used in forests where the
est, trees 2 1 0 cm in diameter are likely to constitute trees make reliable annual rings. All the live trees in
2 9 0 % of vegetation biomass, and would thus suffice a plot are cored, the annual rings they formed over the
MEASURING FOREST NPP

Time: t~
*AGB: 5 6 7

Approach 1 :
Stand Increment = (Z Increments of surviving trees) + (Z Increments(s) of ingrowth)
= ((5.5 - 5.0) + (6.6 - 6.0)) + (2.5 - 2.0)

Approach 2:
Stand Increment = (ZAGB at t, - ZAGB at t,) + (Z Biomass of trees that died in the
interval) - [(Biomass of a minimum size tree) x (number of new trees)]

FIG. 2. The two methods for calculating aboveground biomass increment based on measurement of all trees in a plot at
the beginning and end of an interval. Approach 1 is based on tracking individual surviving trees. Approach 2 is based on
measuring all trees in the stand at each census but also requires measurement of trees (a) that died in the interval and (b)
that recruited past the minimum size during the interval. AGB = aboveground biomass of live trees; AGB of a minimum-
sized tree is set at 2 units.

interval of interest are measured. and these radial in- year and is independent of tree size (Swaine et al. 1987,
crements are then converted to biomass increments us- Hartshorn 1990), the underestimate caused by not ac-
ing the allometric equation. counting for dead trees will average -1-3% of the
In Approach 2 (Fig. 2), the estimated total above- initial aboveground biomass per year of census interval
ground biomass of the stand at the beginning of the (as found by Carey et al. 1994). Within-interval growth
interval is subtracted from the estimated total above- of the trees that died in the interval would augment
ground biomass of the stand at the end of the interval, this error. Similarly, in temperate zone plantations or
with no reference to increments of individual trees. successional forests with high tree mortality due to
This difference then has to be adjusted for both tree thinning. either natural or anthropogenic. the "miss-
mortality and ingrowth. For the mortality correction, ing" biomass from trees that died in the interval could
the biomass of all trees that died during the interval is be very large. Thus, failing to correct for tree mortality
estimated from their initial diameters and added to the when using Approach 2 (e.g., Weaver and Murphy
net change in stand biomass. If the measurement in- 1990) can produce a substantial underestimate of NPF?
terval is long ( > 2 yr), attempts should be made to It will also have a high variance, especially in areas
estimate the within-interval growth by all trees that affected by catastrophic disturbance (hurricanes, land-
died; failure to do this will produce an underestimate slides). and in small study plots. where not accounting
of stand increment, especially when large trees died for the death of one very large tree could result in a
during a many-year interval. To correct for ingrowth, severe underestimate of aboveground increment. For
the number of new trees is multiplied by the biomass many published estimates of aboveground increment
of the minimum-diameter tree being measured (see Ap- from both boreal and tropical forests, it is not possible
proach I), and this product is subtracted from the mor- to judge the reliability of the estimates because the
tality-adjusted stand increment. authors failed to state how they dealt with mortality
Approach 2 is often used to calculate aboveground (see Clark et al. 2001. Gower et al. 2001).
biomass increment from large-scale forest inventory Even when the biomass accounting is carried out
data (e.g., Schulze et al. 1999) or for stands that are correctly. several other sources of error can affect es-
remeasured after a long interval (e.g.. 35 yr, Weaver timates of aboveground biomass increment. One is
and Murphy 1990). Failing to make the corrections for over- or underestimating the biomass of very large trees
tree mortality when using Approach 2 will result in (arbitrarily defined as those of diameters >70 cm). In
underestimating the aboveground biomass increment tropical forests. for example. although densities of such
by an amount at least equal to the initial biomass of trees are low (usually < l o % of stems 2 1 0 cm in di-
trees that died in the interval (see above). In tropical ameter). they can comprise 2 5 5 0 % of the total above-
forests, for example. because stand-level mortality of ground biomass (Brown and Lugo 1992. Brown et al.
trees 2 1 0 cm in diameter is usually 1-3% of stems per 1995, Clark and Clark 1996). If the biomass allometry
DEBORAH A. CLARK ET AL Ecolog~calAppl~cation\
360
Vol. 1 1 . No. 2

equation is based on harvest data that do not cover the cause of the faster radial growth of these stem irreg-
largest tree sizes. the biomass estimates for out-of- ularities (cf. Sheil 1995). Second, with the high species
range big trees may be highly inaccurate (Brown and diversity of tropical forests. it is not feasible to develop
Lugo 1992. Brown et al. 1995). species-level allometries that can be used for estimating
Another potential source of error in estimating stand-level biomass increment. Given the wide range
aboveground biomass increment is the progressive loss of wood densities and tree architectures present in a
of trees' biomass through heartrot and branchfall (self- single tropical forest stand. the development of a single
pruning, wind, and lightning damage). If the allometric biomass allometry equation for such forest should be
equations were based on harvesting a representative based on sampling large numbers of trees of all sizes
sample of trees of all sizes, including those that have and conditions. Nevertheless. despite the limited tree
lost mass due to such factors. the resulting estimate of biomass data from different tropical forests, pooling
biomass increment will need to be "corrected" for the data across species and grouping them by broad
these woody biomass losses due to a question of mass climatic zones produces highly significant regression
balance. The difference in a tree's initial and final bio- equations with >90% of the variation in tree biomass
mass over an interval is the net result of the production explained by diameter alone (Brown 1997). How well
of new biomass and the loss during the interval of both the existing tropical allometry equations predict stand
new and old (previously produced) woody material due biomass has been practically untested. To do this would
to processes such as branchfall, crown damage by require comparing the predicted and measured biomass
storms, and heartrot. When substantial woody mass is of independent samples of trees harvested on site or in
lost by trees during the interval. and when the biomass that forest type. In one recent study (Araujo et al. 1999).
allometry is based on representative trees. including the actual fresh biomass of a harvested Amazonian for-
those with such losses. the true production of new est plot was compared to predictions from 14 allometric
aboveground biomass in the interval will be underes- equations developed elsewhere in Amazonia: although
timated by allometry unless the lost mass is added back some of the equations predicted biomass well, others
to the final estimated biomass. The largest mass loss produced highly erroneous estimates (up to 3 18% high-
in most forests is usually branchfall. Estimates of er than actual biomass). Additional sources of error in
branchfall (wood >1 cm in diameter) in 10 tropical tropical forests are the atypical allometry and growth
forest sites ranged from 0.1 to 2.9 Mg C.ha-l.yr-' patterns of three large woody growth forms: palms.
(Clark et al. 2001). Because branchfall can be highly hemiepiphytes. and lianas. Palms. which can be up to
variable from year to year. continuous measurements 25% of the stems 2 10 cm in diameter in tropical forests
over 5-10 yr would likely be needed for a reliable (e.g.. Lieberman et al. 1985). differ from the other trees
"background" rate of branchfall for this correction. in biomass allometry (Brown 1997), and most of their
Also, branches fallen from standing dead trees would growth is apical. Hemiepiphytes (shrubs and trees) also
need to be identified and excluded from the measure- have distinctive allometries. and they can attain large
ments of branchfall mass. If. however. the allometric sizes and densities in some tropical forests. Similarly.
equations are based on the harvest of less damaged lianas. which can account for 30% of canopy leaf area
trees. adjusting the estimated biomass increment for in tropical forests (Putz 1984). strongly contrast with
these types of mass loss would be incorrect. trees in both their biomass allometry (Putz 1983) and
With long intercensus periods (e.g.. 13 yr. Lieberman the way they grow (principally by stem elongation:
et al. 1990: 35 yr. Weaver and Murphy 1990). above- Putz 1990). Applying tree biomass allometric equations
ground increment will be underestimated if ingrowth to these growth forms could produce erroneous esti-
is not considered. Small trees may have grown past the mates of biomass increment. Also, lianas and hemi-
minimum diameter and then died before the remea- epiphytes are missed when their stems do not descend
surement. The impact of missing such trees will depend to near the ground within the plot. The degree to which
on stand structure. In large-stature forest. the ingrowth these growth forms distort estimates of tropical forest
of small trees may have a negligible effect on estimates biomass increment is unknown.
of aboveground increment. Special consideration is also needed for forests that
Several additional factors affect the estimation of are completely deciduous. such as many temperate and
aboveground increment in tropical forests. The height boreal hardwood forests. For these. the aboveground
of diameter measurement should be above the buttress- biomass increment should be estimated from biomass
es and other bole irregularities that commonly occur allometric equations that exclude foliage mass. This
on large tropical trees at breast height (dbh = 1.30- will eliminate the possibility of including the same fo-
1.37 m above the ground). The allometric equations liage twice in NPP*. by estimating it both from leaf
(cf. Brown 1997) relate tree biomass to the cylindrical litterfall and as included in aboveground increment.
bole diameter. which is much smaller than the diameter Aboveground losses.-In addition to the accumula-
at breast height on buttressed trees. Increments based tion of new biomass by plants, some of the organic
on around-buttress measurements will be inflated. both matter produced in an interval is shed or otherwise lost
because of misapplication of the allometry. and be- by plants during the interval. Such losses include
April 200 1 MEASURING FOREST NPP 36 1

leaves, flowers, fruits. material lost to herbivores, bio- such as an extreme drought that causes anomalous leaf
genic volatile organic compounds. and leached organ- drop and a net loss of live canopy foliage. leaf litterfall
ics (Fig. lb). These materials are not considered in could include inappropriate material that would inflate
measurements of biomass increment and need to be the estimate of current-year foliar NPP. Such an oc-
added to it for an estimate of aboveground NPP. currence could invalidate field estimates of current-
1. Fine litterfall.-Although any net increase in live year NPP or of interannual variation in NPP. To develop
leaf biomass is accounted for as part of the above- reliable estimates of fine litterfall. this material should
ground biomass increment (except in strictly deciduous be measured over several years.
forest [see Abo\%eground increments a n d losses: Above- The timing and frequency required for litterfall col-
ground biornass incrernent]). a large additional fraction lection will depend on several factors. Although lit-
of current-interval leaf production, along with other terfall may be very reduced during the leaf-off season
short-lived plant material such as flowers and twigs. in some deciduous forests. fine litterfall estimates
will be shed before the end of a measurement interval. should be based on continuous collection through the
Fine litterfall is one of the largest components of NPP* year. In many forest types. high rainfall and tempera-
and is the most frequently measured in many types of tures during at least part of the year make it necessary
forest. Relatively straightforward to measure directly. either ( I ) to collect litterfall at least every two weeks
fine litterfall nevertheless presents special challenges. to avoid significant losses due to decomposition (cf.
A nontrivial issue is which material to collect. While Stocker et al. 1995). or (2) to measure decomposition
the goal is clear. to quantify all the current interval's rates and back-correct the litterfall biomass according-
new organic matter that is shed by plants aboveground ly. In wet tropical forest in New Guinea, for example.
(Fig. Ib). doing this correctly is not as simple as it Edwards (1977) estimated that twigs < 1 cm in diameter
appears. Clearly appropriate components are fallen lost 36-40% of dry mass before falling into the litter
leaves and plant reproductive parts. Because large fall- traps. Similarly. Frangi and Lugo (1985) suspended
en branches are mostly composed of wood produced large. old leaves from palms in Puerto Rico and found
in years prior to the measurement interval. they should about half their mass was lost through decomposition
be excluded from fine litterfall (but see Aboveground before they fell.
biornass increment). Even small fallen twigs are a mix In all forests. the estimation of fine litterfall can be
of organic matter fixed in the current year and in pre- highly uncertain due to the use of insufficient numbers
vious years. An upper size limit must be set on the of traps to deal with the high intertrap variance in col-
woody material considered as part of fine litterfall. A lected material. Field studies should be based on pre-
reasonable upper bound is a diameter of 1 cm. (At a sampling to quantify this variance and to define the
diameter growth rate of 2 mm/yr. most cross-sectional necessary number of traps. Secondly. confidence in-
area (64%) of a 0.5 cm diameter twig will be from the tervals should be given for any reported litter data.
current year's growth.) In tropical forests. "fine litterfall" often includes
Unfortunately. as documented for tropical forests by large items such as 10-m-long palm leaves. Nonstan-
Proctor (1983). NPP studies have been notoriously non- dard approaches to littertrap design and management
standardized with regard to woody "fine litterfall". In are needed to sample these adequately: otherwise. fine
numerous studies. all size ranges of woody material. litterfall can be severely underestimated (Villela and
including the largest branches. have been counted as Proctor 1999). A second, unquantified cause of litterfall
litterfall. and in others a variety of maximum diameters underestimation in tropical forests is that leaves and
have been used (Clark et al. 2001). This problem is other fine litter are often trapped in the crowns of other
compounded by the fact that authors frequently report plants (palms, understory shrubs), where they are con-
litterfall values without specifying what material was sumed or decompose. As a result, this component of
collected. litterfall is never collected in littertraps.
A second issue is leaf longevity. To the degree that 2. Aboveground 1o.r.re.rto consumers.-Aboveground
leaves live > I yr, annual leaf litterfall will include production can be subject to substantial losses to con-
production from both the current and past years. In sumers. at least in some forest types and at some times
most cases, all this material should be included in NPP* (e.g.. insect outbreaks). Not accounting for leaf her-
due to mass balance considerations. The difference be- bivory and seed and fruit predation may result in un-
tween a tree's live foliage mass at the beginning and derestimates of NPP. Additional production is also lost
at the end of an interval is the net result of new foliage to sap-sucking insects and nectarivores.
production during the interval, and loss during the in- Of these classes of consumption. only herbivory has
terval of both new and old (previously produced) been quantified in forest NPP studies, and only in a
leaves. To account for the total new foliage production small proportion of them. Several studies (cited in
during the measurement interval, the "old" and "new" Lowman 1995) have shown that grazers can consume
leaves in fine litterfall need to be added to the net live as much as 12-30% of leaf biomass. When herbivory
foliage increment (the foliar portion of aboveground is estimated from the percentage of the area missing
biomass increment). However, under some conditions, from leaves collected in littertraps (e.g., Odum and
DEBORAH A. CLARK ET AL. Ecological Application\
362
Vol. 1 I . No. 2

Ruiz-Reyes 1970, Edwards 1977), it is underestimated, 1999) and may be responding strongly to the global
because this method misses the losses of entire leaves temperature increase.
to herbivores. When this method was used concurrently A second component of aboveground NPP* (Fig. 1b)
with studies of individually tracked leaves (a subtrop- that remains unquantified for forests is the leaching of
ical Australian forest [Lowman 19841. a Mexican trop- organics from aboveground plant parts. Studies of cul-
ical dry forest [Filip et al. 19951). the estimated losses tivars suggest that such losses in throughfall may not
based on instantaneous area measurements were only be negligible (0.8 Mg.ha-'.yr-I of organic leachates in
about half those based on tracking leaves (Australia: 8 apple orchards [Dalbro 19551: 6% of carbohydrates
vs. 15% area lost per year: Mexico: 8 vs. 17% area lost leached from young bean leaves in 24 h [Tukey and
per year). Thus. leaf herbivory should ideally be as- Mecklenburg 19641).
sessed on site by the latter method, and the estimated
Belowground increments a n d losses
loss should be used to back-correct the leaf components
of fine litterfall and aboveground biomass increment. Belowground production in forests remains poorly
Similarly, predispersal consumption of fruits and understood due to method challenges and incomplete
seeds reduces the biomass of these materials before measurements. There are seven components of below-
they reach littertraps. and animals can also remove ground NPP* in forests (Fig. lb): biomass increments
these materials from the traps. Not accounting for these in coarse roots and in fine roots. mortality losses of
losses may produce an underestimate of NPP. For ex- new coarse root and fine root production, losses of root
ample. fruit production of the dominant plant species material to belowground consumers. rhizodeposition
in a Puerto Rican palm forest was estimated in a sep- (root exudates. sloughed root cells. etc.). and the car-
arate study to be 14 times the fruit fall measured in bohydrate inputs to mycorrhizal fungi and root nodule
litter traps (Lugo and Frangi 1993). Janzen and Vaz- symbionts. There are no straightforward methods for
quez-Yanes (1991) have estimated that. for nearly all field measurement of any of these components. How-
tropical trees. >50% of seeds are lost to animal con- ever, there are some promising new directions for quan-
tifying belowground NPF?
sumers or fungi. In temperate and boreal forests. pre-
A first need is to define "coarse" and "fine roots."
collection losses of reproductive material may be con-
Fine roots are considered the most biologically active
siderably less, but this needs to be demonstrated. Study
and show rapid turnover while usually contributing lit-
of the relationship between actual production of seeds,
tle to total root biomass in old-growth forest (Vogt et
fruits, and flowers, and collections from litter traps is
al. 1996, Cairns et al. 1997). They are thought to ac-
needed in all forest types, so that NPP* can be corrected
count for a large portion of total annual losses of or-
appropriately.
ganic material in most ecosystems. Large roots, in con-
3. Biogenic b lo la tile organic compounds a n d leached
trast, are thought to turn over slowly, and their con-
organics.-It is not yet known what proportion of for-
tribution to NPP is largely in terms of biomass incre-
est NPP are biogenic volatile organic compounds
ment. The operational distinction between coarse and
(BVOCs), compounds with important effects on at- fine roots varies across studies, and it will be highly
mospheric chemistry (Crutzen et al. 1999). Using the desirable to standardize this. Regardless of the limit
available emissions data and estimates of biome-spe- used, however, the critical point for NPP studies is to
cific factors such as foliar density and climatic variation cover the entire size range of roots.
in emissions. Guenther et al. (1995) modeled global Belowground bi0rna.r.r increment.-
BVOC fluxes. Their model projected that "tropical 1. Net coarse root increment.-For direct field mea-
woodlands" (rain forest, seasonal, drought-deciduous, surement of this NPP* component, two kinds of in-
and savanna) produce nearly half of total world BVOC formation are needed: the standing biomass and size
emissions. Combined emissions of isoprene, monoter- distribution of coarse roots in the study plot, and their
penes. and other reactive volatile organic compounds size-dependent increments. Developing reliable esti-
were estimated to be 0.31. 0.15. and 0.21 mates for either of these is problematic in any forest.
Mg C.ha-'.yr-' for tropical rain forests. tropical mon- The best estimates of coarse root biomass will come
tane forests. and tropical seasonal forests. respectively. from a two-pronged effort (Bledsoe et al. 1999) that
For boreal conifer forest. in contrast. they estimated combines: ( 1 ) sampling of coarse roots in replicated
combined emissions to be <0.02 Mg C.ha-l.yr-l. These monoliths in the areas away from tree stems, and (2)
estimates suggest that BVOCs are an insignificant pro- a biomass allometry approach based on excavation and
portion of forest NPP. However. Guenther et al. (1995) harvesting individual trees' coarse root systems within
underlined the uncertainties in these first-order calcu- a given radius of the stem and relating them to tree
lations due to the few data and limited understanding diameter. Then, the tree diameter distribution in a study
of underlying processes. They estimated the uncertain- plot and the monolith data can be combined to estimate
ty around the tropical emissions to exceed a factor of the standing crop of coarse roots. Backhoes and fire
three. Furthermore. BVOC emissions increase with trucks have been used to excavate the coarse roots of
temperature (Lerdau and Keller 1997. Constable et al. trees in plantations and forest in the temperate and
April 2001 MEASURING FOREST NPP 363

boreal regions (Bledsoe et al. 1999). In tropical forests. surrogate for annual production, given the potential for
where root systems have been excavated for only a year-round belowground production and mortality. Se-
handful of individual trees (see Clark et al. 2001), re- quential coring methods based on simply summing sta-
liable estimates of the allometry between coarse root tistically significant increments in fine-root biomass be-
biomass and aboveground tree biomass await a greatly tween sampling periods would lead to underestimates
expanded database: this will likely be built up slowly in those forests where root production and mortality
and opportunistically (relying on recently cut soil pro- co-occur (even if total fine root stocks are unchanging).
files along roadsides). Such errors are compounded when temporal patterns
Techniques still need to be worked out for directly of fine root production and mortality differ among soil
measuring radial and longitudinal growth rates of the depths, as found in northern hardwood forest by Hen-
different sizes of coarse roots. To date, stand-level drick and Pregitzer (1996).
coarse root increment has been estimated based on the In spite of the challenges, assessment of fine root
diameter distributions and increments of the trees and dynamics should receive high priority, given the po-
allometric relations between root biomass and above- tentially large fraction of NPP that can be involved (cf.
ground tree biomass. A potential problem with this an estimated 32-49% in two classes of Alaskan taiga
indirect method is that the smallest roots commonly forests [Ruess et al. 19961). Because all currently avail-
classified as "coarse roots" (often 2-10 mm in di- able methods have problems associated with them, cur-
ameter) are unlikely to be well characterized in biomass rent concensus (see Vogt et al. 1998, Fahey et al. 1999)
allometric approaches because of breakage and incom- is that the most robust direct estimates will come from
plete sampling in excavations of root systems; these using multiple techniques in parallel, and cross-check-
small, coarse roots may be considerably more dynamic ing results between them. The most promising methods
(i.e.. account for more NPP*) than larger roots. combination appears to be (1) the compartment model
2. Net j n e root increment.-This NPP* component approach to sequential coring (Fairley and Alexander
is any net increase in live fine root biomass during the 1985. Santantonio and Grace 1987. Publicover and
measurement interval. To measure this fraction of be- Vogt 1993). and (2) minirhizotrons (cf. Aerts et al.
lowground production would require quantifying initial 1989, Steele et al. 1997. Fahey et al. 1999). The com-
and final biomass of live fine roots. Methods for as- partment model method combines repeated destructive
sessing standing stocks of fine roots are discussed be- sampling to assess standing stocks of live and dead fine
low. roots, with in situ measurement of fine root decom-
Belowground losses.- position under differing environmental conditions
1. Dead coarse roots.-The same mass balance con- (e.g.. soil moisture, temperature). A simple flow model
siderations that affect estimates of foliar NPP apply to can then be used to estimate the biomass of fine roots
estimating the production of coarse roots: both coarse produced and lost during the intervals between sam-
root increment (see Net coarse root increment) and ples, based on the mass of live and dead roots in suc-
coarse root mortality need to be quantified. The tech- cessive samples and on the climatic conditions for de-
niques needed to measure standing stocks of dead composition in each interval. The minirhizotron ap-
coarse roots are the same as those needed for assessing proach, based on sequential video images taken from
live coarse root biomass. To estimate coarse root mor- within a buried transparent tube, enables direct obser-
tality over a measurement interval (Fig. l b ) requires vation of the production and mortality of fine roots, at
estimating the net change in the stocks of dead coarse differing depths in the soil profile. However, this tech-
roots; decomposition of these larger roots is unlikely nique is extremely labor intensive and may be hard to
to be a big source of error. replicate sufficiently to deal with the spatial variability
2. D e a d j n e roots.-Fine root losses and even fine in fine root biomass. It involves soil disturbances on
root standing stocks are difficult to measure in forests installation that could affect root behavior (although
(see Bledsoe et al. 1999, Fahey et al. 1999). Estimates intense biotic soil-mixing in some forests likely re-
of standing stocks of fine roots are highly uncertain stores soil structure within a few months). Using mini-
due to the notorious temporal and spatial variability in rhizotrons to estimate fine root production also requires
fine root biomass (e.g.. Carvalheiro and Nepstad 1996, data on fine root biomass. The data on root growth and
Ostertag 1998). Second, particularly in strongly sea- mortality are based on root length, and relationships
sonal tropical forests, live roots can occur much deeper between length and biomass are needed to convert mea-
in the soil profile than has been assumed (e.g., to 18 surements to biomass. The demographic data provided
m depth in an eastern Amazonian forest [Nepstad et by minirhizotrons can be critical. however, because er-
al. 1994: see also Trumbore et al. 19951). These factors. roneous estimates of fine root life-span will produce
along with the challenge of distinguishing live from large errors in estimates of fine root production. In the
dead fine roots. make it difficult to assess fine root absence of direct data. some past estimates of below-
dynamics. Techniques that sample only the top of the ground carbon cycling (cf. Trumbore et al. 1995) have
soil profile can underestimate fine root production and been based on the assumption that fine roots turn over
mortality. Peak fine root biomass cannot be used as a annually. A number of recent minirhizotron studies
DEBORAH A. CLARK ET AL. Ecological Applications
364
Vol. 1 I . No. 2

have shown. however, that tree fine roots can have a high upper bound for total belowground NPP under
much shorter life-spans (e.g.. 85 d in an oak forest in steady-state conditions. When concurrent data on an-
Spain. Lopez et al. 1998: 15-180 d in 10 of 14 studies nual litterfall and annual soil respiration are available,
of temperate forest trees reviewed by Eissenstat and this method should be used to chkck the plausibility of
Yanai 1997: <14 d for >40% of Prunus aviurn roots. field estimates of belowground components of NPP*
Black et al. 1998): with such short life-spans. fine roots (cf. Nadelhoffer et al. 1998). Multiple issues arise with
would account for considerably more NPP than would respect to interpretation, however.
be estimated based on a 1-yr life-span. In contrast. One important uncertainty is the magnitude of root
where mean fine root longevities are found to be much respiration. Some have assumed it to account for -50%
greater than 1 yr, assuming a 1-yr life-span would result of TBCA (cf. Raich et al. 1991). In two recent studies
in overestimating the carbon allocation to fine root pro- in temperate deciduous forests, however, root respi-
duction. ration was estimated to be 33% (Bowden et al. 1993)
3. Rhizodeposition, inputs to symbionts, a n d root and 90% (Thierron and Laudelout 1996) of soil res-
losses to consumers.-Recent studies have indicated piration. Using the C-balance method in eastern Ama-
that these three belowground processes may accohnt zonia. Trumbore et al. (1995) estimated root respiration
for a large proportion of total NPP in forests. Rhizo- to be 50-65% of soil respiration. This estimate was
deposition (soluble root exudates, mucilaginous ma- based on the assumption that the carbon in the standing
terial, sloughed root cells. etc. [Darrah 19961) and ex- stocks of fine roots (<1 mm in diameter) equaled the
port to mycorrhizae were estimated to represent >30% total belowground carbon input by plants to the soil
of NPP for ponderosa pine seedlings (Norton et al. over an annual cycle. Including faster fine root turn-
1990). In a mature fir stand. carbon allocation to my- over, losses of roots 2 1 mm in diameter, or rhizode-
corrhizae was estimated at 15% of forest NPP (Vogt et position would have decreased their estimate of root
al. 1982). Bekku et al. (1997) found root exudates to respiration. Unfortunately, estimating forest-level root
account for 3-13% of NPP for temperate weed seed- respiration by more direct methods is made difficult by
lings. Rhizodeposition is as yet unquantified at the for- potential measurement artifacts (cf. Clinton and Vose
est level. Similarly unknown is the amount of new root 1999) and the high spatio-temporal variation of root
biomass that is consumed by soil organisms. Eissenstat biomass.
and Yanai (1997) have recently argued that root her- A second issue arising with the C-balance method
bivory is likely to be much greater than is generally is its requirement that SOC be close to steady state
appreciated. given the lack of structural defenses in (Nadelhoffer et al. 1998). This assumption may be vi-
young roots. olated in today's forests, given that forest soils contain
Constraining belowground NPP.-Given the chal- large quantities of SOC (Dixon et al. 1994. Schlesinger
lenges for measuring the components of belowground 1997) that can be very dynamic (cf. Davidson and
NPP*. an indirect method based on carbon mass bal- Trumbore 1995. Binkley and Resh 1999), and given
ance (Raich and Nadelhoffer 1989) can be used to con- that SOC turnover increases with temperature (Trum-
strain estimates of total belowground NPP. An upper bore et al. 1996. Goulden et al. 1998. Lindroth et al.
bound for total BNPP can be estimated by subtracting 1998), and that global temperatures are increasing. Re-
the carbon input to the soil by aboveground litterfall sulting errors. however. will be conservative when
from that emitted in soil respiration. Total soil respi- TBCA is taken as an upper bound for BNPP. because
ration comprises: (1) root respiration. (2) respiration efflux of old SOC in soil respiration will inflate the
by microbes decomposing the litter and other materials TBCA estimate. Two other potential non-steady-state
shed by plants both above- and belowground. and (3) processes. however. would have the opposite effect: net
a minor level of respiration by other soil organisms. If C losses from the soil profile as dissolved organic car-
soil organic carbon (SOC) is close to steady state. if bon. or any net accumulation of SOC. If either of these
fine root biomass and coarse root biomass are also close occur at nontrivial rates. they would need to be factored
to steady state, and if fine litterfall is the only signif- into the C-balance calculation.
icant aboveground source of carbon, then subtracting Given the potential impacts of year-to-year climatic
the litterfall carbon from the carbon respired from the variation and of local to large-scale disturbances on
soil provides an estimate of total belowground carbon forest production (cf. Schulze et al. 1999), root dy-
allocation by the plants (TBCA). TBCA is the summed namics are likely to vary among years. Applying the
carbon in root respiration plus all belowground com- C-balance method in forest NPP studies will therefore
ponents of NPP (Fig. l a ) . Neglected in this calculation. be most useful when based on multiple years of con-
however. are several potential aboveground carbon current litterfall and soil respiration measurements that
sources in forests: decomposing coarse woody debris. are well replicated within the study area. thus averaging
leached organics from aboveground vegetation. dead out the spatiotemporal variation in these two fluxes.
animals and frass. and carbon in dry and wet deposi- Finally, when estimated TBCA is compared across sites
tion. If any of these is significant, their omission will and studies, careful attention needs to be given to the
inflate estimated TBCA. The estimate of TBCA is thus litterfall methods in each study. The great variation in
April 2001 MEASURING FOREST NPP 365

litterfall definitions and methods in past NPP studies total NPP (lower bound) to be 11.8 Mg C.ha-'.yr-I,
(see Aboveground losses: Fine litterfall) could produce based on the existing field data (aboveground biomass
spurious inter-site differences in BNPP as estimated by and aboveground biomass increment [Kira et al. 19671)
the C-balance method. and additional NPP components estimated by Clark et
al. (2001). We emphasize that this exercise is heuristic
Neb$, Stored Nonstruc~turalCarbohydrates in intent. As detailed in Table 1, it involves many ar-
Some net primary production can be allocated to bitrary assumptions and builds on fragmentary data
storage by trees both above- and belowground in the from other studies. It does, however, provide insights
form of nonstructural carbohydrates. As noted by War- into the uncertainty associated with existing field meth-
ing and Schlesinger (1985), any significant stand-level ods for estimating NPP.
changes from year to year in the amount of such storage The first conclusion that can be made from this ex-
will provide problems for the estimation of total NPP. ercise (Table 1) is that, at any study site, some of the
This issue has not yet been addressed in field studies methodological issues will be particular to the forest
of forest NPP. type being studied (e.g., the potential for double-ac-
counting foliage production in deciduous forests, large
palm leaves skewing litter estimates in tropical forests,
In addition to the lack of direct measurements of and the large contribution of NPP from ground cover
many NPP* components, plot biases and a lack of rep- or the shrub layer in some boreal forests). Thus, field
lication in space and time decrease the reliability of methods need to be carefully adjusted to the conditions
existing estimates of forest NPP. Within the boreal re- of the site; failing to do this can produce major errors
gion, the fire-disturbance cycle (Price and Apps 1996, in estimated NPP. Second, the relative impact of some
Schulze et al. 1999) has pervasive effects on forest of the more general methods issues can vary among
carbon dynamics. Tropical moist forests have strong forest types. For example. precollection decomposition
spatial variation in forest structure and edaphic con- is likely to most strongly affect fine litterfall estimates
ditions at the local scale (cf. Richter and Babbar 1991, in the warm, humid conditions of tropical moist and
Clark et al. 1998, Laurance et al. 1999), and are likely wet forest. In forests with higher root:shoot ratios, be-
to show corresponding within-forest variation in pro- lowground methods issues may have higher relative
ductivity. For example, Gower ( 1987) found a doubling impacts on NPP estimates. Third, and most important,
of fine root biomass between two soil types in one small most of the procedural problems in NPP field studies
area of lowland forest in Costa Rica. The small study result in underestimating NPP. Some of these can in-
plots generally used for ecological studies in these for- dividually produce large underestimates of NPP.
ests often have an over-representation of large trees Summed, the error estimates in Table 1 suggest the
(Brown and Lugo 1992, Brown 1997) and show high potential for underestimating NPP by >200% in this
interplot variance in total aboveground biomass due to forest. Neglecting NPP* components accounts for most
the spotty distributions of the large trees (Brown et al. of this potential underestimation, followed by errors
1995). To assess NPP in such internally heterogeneous associated with measuring aboveground increment and
forests, measurements should be replicated in space, fine litterfall.
ideally in a stratified random fashion with respect to Most forest NPP studies involve an important num-
the major gradients of within-forest variation. Simi- ber of these errors and omissions. For example, in their
larly, NPP:p components in a given forest can show recent review of field NPP* data in boreal forests, Gow-
substantial temporal changes due to interannual cli- er et al. (2001) found that even the most complete
matic variation, as has been found for aboveground studies failed to assess carbon allocation to mycorrhi-
biomass increment in boreal forest (Yarie 1997) and zae. and that fine root production and foliar herbivory
for tree growth rates in tropical wet forest (Clark and were usually also not measured. Further, no study in-
Clark 1994). For a robust assessment of NPP in a given cluded estimates of root exudates, root herbivory,
forest, measurements should be made over multiple BVOC production, losses of organic leachates, seed
yearly cycles. consumption, or losses to sap-suckers. With the ex-
ception of fine root production, this same list of NPP*
OF NPP
UNDERESTIMATION IN FIELDSTUDIES components as well as fine litterfall were all unmea-
Given the uncertainties and the incomplete mea- sured in the classic study of 40-yr-old Douglas fir
surements characterizing forest NPP studies to date stands by Keyes and Grier (1981) that focused on how
(see Clark et al. 2001, Gower et al. 2001), we can only ANPP and BNPP change with differing site productiv-
make educated guesses about the impact of frequently ity. Similarly, in our review of the existing field NPP'p
used procedures on estimates of total NPP. To illustrate data for tropical forests, we found that all studies in-
the issues and their potential relative importance, how- volved many of the omissions and methods problems
ever, we have developed error estimates (Table l ) for noted in Table 1 (see Clark et al. 2001). Thus, in ad-
one example site, a tropical moist forest (Khao Chong, dition to the uncertainty around most NPP numbers due
Thailand) for which Clark et al. (2001) have estimated to sampling issues, the field research to date appears
Ecological Applications
366 DEBORAH A. CLARK ET AL
Vol. 11, No. 2

TABLE 1. Estimated potential impact of problems associated with common procedures in field measurements of NPP*
components. for a case study site in tropical moist forest (Khao Chong, Thailand).

Potential impact
on estimated NPP*
Procedure (% of NPP* estimate)
Aboveground increment
Using an inappropriate biomass allometry developed off site -20 to +11%
Applying the tropical moist forest equation (Brown 1997) to a tropical wet forest (Clark and
Clark 2000) produced a 79% increase in estimated aboveground biomass compared to that
from the tropical wet forest equation (Brown 1997). W e assume this level of impact on
estimated biomass (either increase or decrease). and a proportional effect on aboveground
increment.
Not correcting for tree mortality (Approach 2) -25%
W e assume (see text) that annual biomass loss through tree mortality = 2% of aboveground
biomass (= 3.3 Mg C.ha-'.yrrl). This reduces the (correctly) measured aboveground in-
crement (3.0 Mg C.ha-'.yrrl) to 0.
Not accounting for growth of trees that died in the interval -3 to -0.3%
W e assume that trees that die contribute 1% of the aboveground biomass increment per year
during the intercensus interval. given that tree mortality in tropical forests averages 1-3%
(see Aboveground increments a n d losses: Aboveground biomass increment). The estimate
error range is for census intervals of 1-10 yr.
Not accounting for ingrowth
W e assume that the percentage of new stems = the percentage of dead stems (both 2%), that
stem density is 500 trees ( 210-cm diameter) per hectare, and that the biomass increment
of a new tree is 29 kg (the difference between a tree of 12-cm diameter and that of the
minimum-sized tree, at 10-cm diameter), per the tropical moist forest allometric equation
of Brown (1997).
Using an allometry based on harvest data that do not cover the larger tree sizes -3 to +6%
We assume that harvested trees underlying the allometry were all 5 7 0 - c m diameter. that 25%
of total stand biomass (167 Mg Clha; Kira et al. 1967) is in larger trees, that biomass
increment is proportional to biomass. and that the error in projected biomass of the out-of-
range trees can be from -50% to + 100%. depending on the allometric equation used.
Not correcting for branchfall and heartrot
W e assume that the long-term average mass loss through branchfall and heartrot by surviving
trees is 2.0 M g C.ha-l.yr-'. This material needs to be counted as NPP* (a mass balance
correction) when the biomass allometry is based on representative trees; no correction should
be made when the allometry is based on unrepresentative (undamaged) trees.
Applying the tree biomass allometry equation to palms, lianas, and hemiepiphytes
There is currently no basis for estimating the errors due to these life forms (see Aboveground
increments a n d losses: Abovegroztnd biomass increment).
Not measuring the lianas and hemiepiphytic trees and shrubs that do not extend down to ground
level
We assume such stems account for a maximum of 10% of the total aboveground biomass
increment.
Fine litterfall
Not correcting for decomposition before material falls in traps
We assume that small wood and leaves are 5.6 Mg C.harl.yr-' ( 9 5 9 of fine litter), and that
they average a 20% mass loss from decomposition before being collected in traps.
Including large wood (>1-cm diameter)
W e assume that large wood litter ( > I - c m diameter) is 1.0 Mg C,ha-',yr-I.
Not using additional, larger traps to collect large leaves
W e assume that large palm leaf litter that is not sampled by standard litter traps (cf. Villela
and Proctor 1999) can be up to 3.0 Mg C.ha ',yr-I.
Not correcting for leaf herbivory
W e assume that 15% of the mass of new foliage is lost to herbivores: we estimate leaf litter
(4.4 Mg C.ha-'.yr-l) as 75% of total estimated fine litter, and we back-calculate the herbivory
loss from this value.
Not correcting for precollection consumption of seedslfruits
W e assume that 50% of seeds and fruits are consumed before falling (see Abovegroztnd losses:
Abovegroztnd losses to consumers), and that seeds and fruits comprise 5% of the trapped
fine litterfall.
Other NPP* components
Not measuring carbohydrates consumption by sap-suckers -4%
W e assume that the carbon lost to sap-suckers is 5% of the C in new foliage, which we
calculate as leaf litter (0.75 total litter), back-corrected for precollection decomposition
(20%) and herbivory (15%) and missed large palm leaves.
Not measuring emissions of biogenic volatile compounds -8 to -3%
We use the estimate of Guenther et al. 1995 for tropical rain forest total BVOCs, and as an
upper bound, 3 X this value, the (minimum) uncertainty they cite for this value.
Not measuring organics leached from aboveground plant parts -3%
W e use the value of leached organics reported for an apple orchard, and assume C is 50% of
these compounds (see Aboveground losses: Biogenic volutile organic compounds a n d
leached organics).
MEASURING FOREST NPP

TABLE1. Continued

Potential impact
on estimated NPP*
Procedure (% of NPP* estimate)
p~

Not measuring rhizodeposition and C export to nodules or mycorrhizae -30 to 3%

W e use the range of values reported in two studies (but the lower estimate, 3 9 , is for root

exudates alone; see Belo+vground increments a n d losses).

Assuming coarse root increment is proportional to aboveground increment -4 to +4%

W e assume coarse root biomass is 30% aboveground biomass, and calculate potential errors

based on the true ratio of coarse root increment to coarse root biomass being from 50%

less to 50% more than the ratio of aboveground increment to aboveground biomass.

Not accounting for net increases in fine root biomass -7 to 0%


in aggrading forests, or during recovery from disturbance or climatic stress. fine root mass

could increase during the interval. W e assume that initial fine root biomass is 1 % of above-

ground biomass, and that a maximum yearly increase is 50% of initial fine root mass.

Assuming a I-yr lifespan for fine roots 2 5 %


W e assume that fine root life-span is actually four months, and that fine root production was

originally estimated (based on annual turnover) at 50% of the 2.0 Mg C,ha-l.yr-' of total

BNPP originally estimated for the site.

Not correcting for consumption of live roots by soil fauna -,??%

There are no data available for estimating root herbivory in forests.

Note: Negative values indicate underestimation of NPP* (the percentage by which the original NPP* estimate should be
increased); positive values indicate overestimation.

likely to have produced significant underestimates of erotrophic respiration in a forest with sufficient pre-
total forest NPP. This clearly has large implications for cision for validating tower-derived NEE.
regional and global extrapolations of carbon fluxes Nevertheless, there are multiple ways NPP" mea-
based on NPP data from field studies. surements can be used to provide cross-checks and in-
terpretations of tower flux data. For example, a terres-
RELATINGFIELDNPP DATATO WHOLE-FOREST trial carbon sink indicated by a negative NEE should
FLUXMEASUREMENTS be manifested as increments in either vegetation bio-
Major research efforts are currently being directed mass. soil organic carbon, or both. If the eddy covari-
at obtaining estimates of forest-level carbon exchange ance estimate of NEE exceeds the summed biomass
with the atmosphere. using eddy covariance techniques increments (Fig. lb) for that year, such a disagreement
(cf. Goulden et al. 1998, Lindroth et al. 1998, Running would point either to net accumulation of carbon in the
et al. 1999, Schulze et al. 1999). These methods are soil, or to problems with one or more of the methods.
still experimental and incorporate significant uncer- Rough agreement between the summed increments and
tainties (Goulden et al. 1996). Concurrent ground mea- measured NEE, on the other hand, would increase con-
surements at the tower sites can thus be of value for fidence in both methods. Similarly, an increase in fine
providing cross-checks of the flux data (Greco and Bal- root production and/or fine root stocks from one year
docchi 1996). to the next would be encouraging support for greater
There are multiple ways NPP* measurements can be measured nighttime fluxes in the second year. Com-
used in this context. even though they cannot contribute bining on-going measurements of soil respiration (and,
to a direct quantitative test of tower flux data. The eddy ideally, plant respiration [see Lavigne et al. 19971) with
covariance studies measure net ecosystem exchange NPP* studies would further extend the types of com-
(NEE) of CO, with the atmosphere. The relationship parisons that could be made this way with eddy co-
between NPP and this quantity is: variance flux estimates. Given the importance of de-
veloping eddy covariance techniques for better under-
NEE = R, - NPP (1) standing ecosystem carbon exchange. it would seem
where NEE is the net flux of CO, carbon between the vital at this stage to maintain highly quality-controlled
forest and the atmosphere at the site (positive NEE concurrent NPP studies at the flux measurement sites.
indicates net input of C 0 2 carbon to the atmosphere),
AND RECOMMENDATIONS
CONCLUSIONS
and R,, is heterotrophic respiration. Thus, NEE is a
small difference between two large fluxes. As such, it Existing NPP estimates for the world's forests have
cannot be validated quantitatively by field measure- been based on incomplete. and in some cases, inap-
ments. As discussed above, forest NPP* cannot cur- propriate field measurements. As a result, forest NPP
rently be estimated with the needed precision. A change may be significantly underestimated. For some forests,
in just one of the many fluxes that contribute to NPP it is possible that many of the unmeasured NPP* com-
could strongly affect NEE (Frolking et al. 1996). Sec- ponents are trivial and/or that different procedural er-
ondly, it is not currently possible to quantify total het- rors cancel each other out, with the result that existing
368 DEBORAH A . CLARK E T AL. Ecological Applications
Vol. 1 1 , No. 2

field-based N P P e s t i m a t e s a r e c l o s e t o a c c u r a t e . D e - J. E. Hooker. 1998. Differences in root longevity of some


t e r m i n i n g w h e r e a n d if this i s t h e c a s e , h o w e v e r , w i l l tree species. Tree Physiology 18:259-264.
Black, T. A., et al. 1996. Annual cycles of water vapour and
r e q u i r e n e w field s t u d i e s that e n c o m p a s s a l l N P P * c o m - carbon dioxide fluxes in and above a boreal aspen forest.
p o n e n t s (Fig. l b ) , f o r at least a s e t o f b e n c h m a r k sites Global Change Biology 2:219-229.
across the major forest types. Although such studies Bledsoe, C. S., T. J. Fahey, F. P. Day, and R. W. Ruess. 1999.
will r e q u i r e i n t e n s e fieldwork a n d t h e resolution of m a - Measurement of static root parameters: biomass, length,
and distribution in the soil profile. Pages 413-436 in G . P.
jor methods challenges, they should be made a high
Robertson, C. S. Bledsoe, D. C. Coleman, and P. Sollins,
priority. W i t h o u t s u c h a n e f f o r t , g r e a t uncertainties will editors. Soils methods for long-term ecological research.
r e m a i n w i t h r e s p e c t t o t h e m a g n i t u d e o f f o r e s t NPP, Oxford University Press, New York, New York. USA.
h o w it varies a c r o s s f o r e s t types, a n d h o w i t is r e - Bowden, R. D., K. J. Nadelhoffer, R. D. Boone, J. M. Melillo,
s p o n d i n g t o c h a n g i n g a t m o s p h e r i c c o m p o s i t i o n a n d cli- and J. B. Garrison. 1993. Contributions of aboveground
litter. belowground litter, and root respiration to total soil
m a t e . T o r e s o l v e t h e s e uncertainties will r e q u i r e b o t h
respiration in a temperate mixed hardwood forest. Canadian
i m p r o v e d practices w i t h c u r r e n t field m e t h o d s a n d n e w Journal of Forest Research 23:1402-1407.
techniques for measuring processes such as rhizode- Brown, I. F., L. A. Martinelli, W. W. Thomas. M. Z. Moreira,
position a n d B V O C e m i s s i o n a t t h e f o r e s t level. M e a - C. A. Cid Ferreira, and R. A. Victoria. 1995. Uncertainty
surements at each site should b e designed to sample in in the biomass of Amazonian forests: an example from
Rondonia. Brazil. Forest Ecology and Management 75:
a n u n b i a s e d f a s h i o n a c r o s s t h e i m p o r t a n t g r a d i e n t s of
175-189.
spatial variation. a n d s h o u l d b e c o n t i n u e d t h r o u g h m u l - Brown, S. 1997. Estimating biomass and biomass change of
tiple a n n u a l c y c l e s . T w o critical n e e d s will b e d o c u - tropical forests: a primer. Forestry Paper 134, Food and
m e n t a t i o n o f t h e m e t h o d s used, a n d p r e s e n t i n g t h e sta- Agriculture Organization, Rome, Italy.
tistical uncertainty a r o u n d t h e e s t i m a t e s o f N P P * c o m - Brown, S., and A. E. Lugo. 1992. Aboveground biomass
estimates for tropical moist forests of the Brazilian Ama-
p o n e n t s . A s part o f t h e s e efforts, t h e C b a l a n c e m e t h o d zon. Interciencia 17:8-18.
(Raich and Nadelhoffer 1989) should be extended to Cairns, M. A,, S. Brown, E. H. Helmer, and G. A. Baum-
p r o v i d e a first-order e s t i m a t e o f total B N P P a t e a c h gardner. 1997. Root biomass allocation in the world's up-
site; this will r e q u i r e o n - g o i n g m e a s u r e m e n t o f soil land forests. Oecologia 111:l-11.
respiration a n d refined m e t h o d s f o r e s t i m a t i n g s t a n d - Carey, E. V., S. Brown, A. J. R. Gillespie, and A. E. Lugo.
1994. Tree mortality in mature lowland tropical moist and
l e v e l r o o t respiration (cf. R y a n e t al. 1 9 9 6 , C l i n t o n a n d tropical lower montane moist forests of Venezuela. Biotro-
V o s e 1999). D e d i c a t i n g substantial r e s o u r c e s a n d hu- pica 26:255-265.
m a n effort i n a "crash p r o g r a m " o f s u c h distributed Carvalheiro, K. D., and D. C. Nepstad. 1996. Deep soil het-
field s t u d i e s w o u l d g r e a t l y i m p r o v e c u r r e n t u n d e r s t a n d - erogeneity and fine root distribution in forests and pastures
of eastern Amazonia. Plant and Soil 182:279-285.
i n g o f t h e c a r b o n d y n a m i c s of t h e w o r l d ' s forests.
Clark, D. A , , S. Brown, D. W. Kicklighter, J. Q. Chambers,
J. R. Thomlinson, J. Ni, and E. A. Holland. 2001. NPP in
tropical forests: an evaluation and synthesis of existing field
This work resulted from a series of workshops conducted data. Ecological Applications 11:371-384.
at the National Center for Ecological Analysis and Synthesis, Clark, D. A,. and D. B. Clark. 1994. Climate-induced annual
a Center funded by the U.S. National Science Foundation variation in canopy tree growth in a Costa Rican tropical
(Grant Number DEB-9421535), the University of California rain forest. Journal of Ecology 82:865-872.
at Santa Barbara, and the State of California. We thank D. Clark, D. B., and D. A. Clark. 1996. Abundance, growth and
B. Clark for extensive critical input during the writing of this mortality of very large trees in neotropical lowland rain
paper, M . G . Ryan for greatly helping us clarify our concep- forest. Forest Ecology and Management 80:235-244.
tual framework. and M . Lerdau, P. Matson, K. Nadelhoffer, Clark. D. B.. and D. A. Clark. 2000. Landscape-scale vari-
and an anonymous reviewer for constructive comments on ation in forest structure and biomass in a tropical rain forest.
the manuscript. During preparation of this paper, D. A. Clark Forest Ecology and Management 137: 185-198.
was supported by grants from the Department of Energy (DE- Clark. D. B., D. A. Clark, and J. M . Read. 1998. Edaphic
FG02-96ER62289), the National Science Foundation (DEB- variation and the mesoscale distribution of tree species in
9629245), and the Andrew W. Mellon Foundation. a neotropical rain forest. Journal of Ecology 86:101-112.
Clinton. B. D., and J. M . Vose. 1999. Fine root respiration
in mature eastern white pine (Pinzcs srrobzcs) in situ: the
Aerts, R., E Berendse, N. M . Klerk, and C. Bakker. 1989. importance of COZ in controlled environments. Tree Phys-
Root production and root turnover in two dominant species iology 19:475-479.
of wet heathlands. Oecologia 81:374-378. Constable, J. V. H., A. B. Guenther. D. S. Schimel, and R.
Araujo, T. M.. N. Higuchi. and J. Andrade de Carvalho. Jr. K. Monson. 1999. Modelling changes in VOC emission in
1999. Comparison of formulae for biomass content deter- response to climate change in the continental United States.
mination in a tropical rain forest site in the state of Para. Global Change Biology 5:791-806.
Brazil. Forest Ecology and Management 117:43-52. Crutzen. P. J.. R. Fall, 1. Galbally, and W. Lindinger. 1999.
Bekku, Y.. M . Kimura, H. Ikeda. and H. Koizumi. 1997. Parameters for global ecosystem models. Nature 399:535.
Carbon input from plant to soil through root exudation in Dalbro, S. 1955. Leaching of nutrient from apple foliage.
Digitnria ndscendens and Ambrosia artemisiifolia. Ecolog- Pages 770-778 in Proceedings of the 14th International
ical Research 12:305-312. Horticultural Congress. International Horticultural Asso-
Binkley, D.. and S. C. Resh. 1999. Rapid changes in soils ciation, Paris, France.
following Eucnl?prus afforestation in Hawaii. Soil Science Darrah. P. R. 1996. Rhizodeposition under ambient and el-
Society of America 63:222-225. evated CO, levels. Plant and Soil 187:265-275.
Black, K. E., C. G . Harbron, M . Franklin, D. Atkinson, and Davidson, E. A,, and S. E. Trumbore. 1995. Gas diffusivity
April 2001 MEASURING FOREST NPP 369

and production of CO, in deep soils of the eastern Amazon. generation and management. UNESCO and The Parthenon
Tellus 47B:550-565. Publishing Group. Paris, France.
Dixon, R. K., S. Brown, R. A. Houghton, A. M. Solomon, Keyes, M. R., and C. C. Grier. 1981. Above- and below-
M. C. Trexler, and J. Wisniewski. 1994. Carbon pools and ground net production in 40-year-old Douglas-fir stands on
flux of global forest ecosystems. Science 263:185-190. low and high productivity sites. Canadian Journal of Forest
Edwards, P. J. 1977. Studies of mineral cycling in a montane Research 111599-605.
rain forest in New Guinea. 11. Production and disappearance Kira. T., H. Ogawa, K. Yoda, and K. Ogino. 1967. Compar-
of litter. Journal of Ecology 65:971-992. ative ecological studies on three main type9 of forest veg-
Eissenstat, D. M.. and R. D. Yanai. 1997. The ecology of etation in Thailand. IV. Dry matter production, with special
root lifespan. Advances in Ecological Research 27: 1-60. reference to the Khao Chong rain forest. Nature and Life
Fahey, T. J., C. S. Bledsoe, E P. Day, R. Ruess, and A. J. M . in Southeast Asia 5: 149-174.
Smucker. 1999. Fine root production and demography. Laurance. W. F.. P. M . Fearnside. S. G. Laurance, P. Dela-
Pages 543-573 in G . P. Robertson. C. S. Bledsoe. D. C. monica, T. E. Lovejoy. J. M. Rankin-de Merona, J. Q.
Coleman, and P. Sollins, editors. Soils methods for long- Chambers, and C. Gascon. 1999. Relationship between
term ecological research. Oxford University Press, New soils and Amazon forest biomass: a landscape-scale study.
York. New York, USA. Forest Ecology and Management 118: 127-1 38.
Fairley, R. I .. and 1. J. Alexander. 1985. Methods of calcu- Lavigne. M . B., et al. 1997. Comparing nocturnal eddy co-
lating fine root production in forests. Pages 37-42 in A. variance measurements to estimates of ecosystem respi-
H. Fitter. D. Atkinson, D. J. Read, and M. B. Usher, editors. ration made by scaling chamber measurements at six co-
Ecological interactions in soil: plants, microbes. and ani- niferous boreal sites. Journal of Geophysical Research 102:
mals. Blackwell Scientific, Oxford. UK. 28977-28985.
Filip, V., R. Dirzo, J. M . Maass. and J. Sarukhan. 1995. With- Lerdau, M.. and M. Keller. 1997. Controls on isoprene emis-
in- and among-year variation in the levels of herbivory on sion from trees in a subtropical dry forest. Plant. Cell and
the foliage of trees from a Mexican tropical deciduous for- Environment 20:569-578.
est. Biotropica 27:78-86. Lieberman, D.. G . S . Hartahorn. M. Lieberman, and R. Per-
Frangi, J. L., and A. E. Lugo. 1985. Ecosystem dynamics of alta. 1990. Forest dynamics at La Selva Biological Station,
a subtropical floodplain forest. Ecological Monographs 55: 1969-1985. Pages 509-521 in A.H. Gentry, editor. Four
35 1-369. neotropical rainforests. Yale University Press, New Haven.
Frolking, S.. et al. 1996. Modelling temporal variability in Connecticut. USA.
the carbon balance of a sprucelmoss boreal forest. Global Lieberman, M.. D. Lieberman, G. S. Hartshorn, and R. Per-
Change Biology 2:343-366. alta. 1985. Small-scale altitudinal variation in lowland wet
Goulden, M. L.. J. W. Munger, S. Fan, B. C. Daube. and S. tropical forest vegetation. Journal of Ecology 73:505-5 16.
C. Wofsy. 1996. Measurements of carbon sequestration by Lindroth, A,. A. Grelle, and A.-S. Moren. 1998. Long-term
long-term eddy covariance: methods and a critical evalu- measurements of boreal forest carbon balance reveal large
ation of accuracy. Global Change Biology 2: 169-182. temperature sensitivity. Global Change Biology 4:443-
Goulden, M. L.. et al. 1998. Sensitivity of boreal forest car- 450.
bon balance to soil thaw. Science 279:214-217. Long. S. P., and P. R. Hutchin. 1991. Primary production in
Gower. S. T. 1987. Relations between mineral nutrient avail- grasslands and coniferous forests with climate change: an
overview. Ecological Applications 1: 139-156.
ability and fine root biomass in two Costa Rican tropical
wet forests: a hypothesis. Biotropica 19:171-175. Lopez. B.. S. Sabate, and C. Gracia. 1998. Fine roots dy-
namics in a Mediterranean forest: effects of drought and
Gower, S. T., 0. Krankina, R. J. Olson. M. Apps, S. Linder,
atem density. Tree Physiology 18:601-606.
and C. Wang. 2001. Net primary production and carbon
Lowman. M . D. 1984. Assessment of techniques for mea-
allocation patterns of boreal forest ecosystems. Ecological
suring herbivory:Is rainforest defoliation more intense than
Applications 11, in prcJss.
we thought?. Biotropica 16:264-268.
Gower, S. T., C. J. Kucharik. and J. M . Norman. 1999. Direct
Lowman. M. D. 1995. Herbivory as a canopy process in rain
and indirect estimation of leaf area index, f,,,,,and net pri-
forest trees. Pages 43 1-455 in M . D. Lowman and N. M.
mary production of terrestrial ecosystems. EOS Remote Nadkarni, editors. Forest canopies. Academic Press, New
Sensing of the Environment 70:29-5 1. York, New York, USA.
Greco, S., and D. D. Baldocchi. 1996. Seasonal variations Lugo. A. E., and J. L. Frangi. 1993. Fruit fall in the Luquillo
of CO2 and water vapour exchange rates over a temperate Experimental Forest. Puerto Rico. Biotropica 25:73-84.
deciduous forest. Global Change Biology 2: 183-1 97. Melillo. J. M., I. C. Prentice. G. D. Farquhar, E.-D. Schulze.
Grier. C. C.. K. M. Lee, and R. M. Archibald. 1984. Effects and 0 . E. Sala. 1996. Terrestrial biotic responses to en-
of urea fertilization on allometeric relations in young Doug- vironmental change and feedbacks to climate. Pages 445-
las-fir trees. Canadian Journal of Forest Research 14:900- 481 in J. T. Houghton. L. G. Meira Filho, B. A. Callander.
904. N. Harris, A. Kattenberg, and K. Maskell. editors. Climate
Guenther, A. B.. et al. 1995. A global model of natural vol- change 1995: the science of climate change. Cambridge
atile organic compound emissions. Journal of Geophysical University Press, London, UK.
Research 100(D):8873-8892. Murphy, P. G.. and A. E. Lugo. 1986. Structure and biomass
Hartshorn, G. S. 1990. An overview of neotropical forest of a subtropical dry forest in Puerto Rico. Biotropica 18:
dynamics. Pages 585-599 in A. H. Gentry, editor. Four 89-96.
neotropical rainforests. Yale University Press. New Haven. Nadelhoffer. K. J.. J. W. Raich, and J. D. Aber. 1998. A global
Connecticut. USA. trend in belowground carbon allocation: comment. Ecology
Hendrick. R. L., and K. S. Pregitzer. 1996. Temporal and 79: 1822-1825.
depth-related patterns of fine root dynamics in northern Nepstad, D. C., et al. 1994. The role of deep roots in the
hardwood forests. Journal of Ecology 84: 167-1 76. hydrological and carbon cycles of Amazonian forests and
Janzen. D. H., and C. Vazquez-Yanes. 1991. Aspects of trop- pastures. Nature 372:666-669.
ical seed ecology of relevance to management of tropical Norton. J. M., J. L. Smith, and M. K. Firestone. 1990. Carbon
forested wildlands. Pages 137-157 in A. Gomez-Pompa, flow in the rhizosphere of Ponderosa pine seedlings. Soil
T. C. Whitmore, and M. Hadley. editors. Rain forest re- Biology and Biochemistry 22:449-455.
DEBORAH A. CLARK ET AL. Ecological Application\
Vol. 1 1 . No. 2

Odum, H . T., and J. Ruiz-Reyes. 1970. Holes in leaves and data: a compartment-flow model. Canadian Journal of For-
the grazing control mechanism. Pages 169-180 in H. T. est Research 17:900-908.
Odum and R. F. Pigeon, editors. A tropical rain forest: a Schlesinger, W. H. 1997. Biogeochemistry: an analysis of
study of Irradiation and ecology at El Verde. Puerto Rico. global change. Second edition. Academic Press, San Diego,
U.S. Atomic Energy Commission. Washington, D.C.. USA. California, USA.
Ostertag, R. 1998. Belowground effects of canopy gaps in a Schulze. E.-D.. et al. 1999. Productivity of forests in the
tropical wet forest. Ecology 79: 1294-1 304. Eurosiberian boreal region and their ~ o t e n t i a lto act as a
Petit. J. R.. et al. 1999. Climate and atmospheric history of carbon ?ink-a synthisis. Global Ch'ange Biology 5(6):
the past 420,000 years from the Vostok ice core, Antarctica. 703-722.
Nature 399:429-436. Sheil. D . 1995. A critique of permanent plot methods and
Price, D. T.. and M. J. Apps. 1996. Summary. Pages 415- analysis with examples from Budongo Forest, Uganda. For-
425 in M . J. Apps and D. T. Price, editors. Forest ecosys- est Ecology and Management 77: 11-34.
tems, forest management and the global carbon cycle. Steele. S. J . , S. T. Gower, J. G. Vogel, and J. M. Norman.
Springer-Verlag, Berlin, Germany. 1997. Root mass, net primary production and turnover in
Proctor, J. 1983. Tropical forest litterfall. I. Problems of data aspen, jack pine and black spruce forests In Saskatchewan
comparison. Pages 267-273 in S. L. Sutton, T. C. Whit- and Manitoba, Canada. Tree Physiology 17:577-587.
more. and A. C. Chadwick, editors. Tropical rain forest: Stocker, G . C., W. A. Thompson, A . K. Irvine, J . D. Fitzsimon.
ecology and management. Blackwell Scientific, Oxford. and P. R. Thomas. 1995. Annual patterns of litterfall in a
UK. lowland and tableland rainforest in tropical Australia. Bio-
Publicover, D. A., and K. A. Vogt. 1993. A comparison of tropica 27:412-420.
methods for estimating forest fine root production with re- Swaine. M. D., D. Lieberman. and F. E. Putz. 1987. The
spect to sources of error. Canadian Journal of Forest Re- dynamics of tree populations in tropical forest: a review.
Journal of Tropical Ecology 3:359-366.
search 23: 1 179-1 186.
Thierron. V., and H. Laudelout. 1996. Contribution of root
Putz, E. 1983. Liana biomass and leaf area of a "Tierra
respiration to total CO, efflux from the soil of a deciduous
Firme" forest in the Rio Negro basin, Venezuela. Biotro-
forest. Canadian Journal of Forest Research 26: 1 142-1 148.
pica 15: 185-189.
Trumbore, S. E., 0. A. Chadwick, and R. Amundson. 1996.
Putz. F. E. 1984. The natural history of lianas on Barro Col-
Rapid exchange between soil carbon and atmospheric car-
orado Island. Panama. Ecology 65: 17 13-1724. bon dioxide driven by temperature change. Science 272:
Putz. F. E. 1990. Liana stem diameter growth and mortality 393-396.
rates on Barro Colorado Island, Panama. Biotropica 22: Trumbore. S. E., E. A. Davidson, P. Barbosa de Camargo, D.
103-105. C. Nepstad. and L . A. Martinelli. 1995. Belowground cy-
Raich, J. W., and K. J. Nadelhoffer. 1989. Belowground car- cling of carbon in forests and pastures of Eastern Ama-
bon allocation in forest ecosystems: global trends. Ecology zonia. Global Biogeochemical Cycles 9 : s 15-528.
70: 1346-1354. Tukey. H. B.. Jr., and R. A. Mecklenburg. 1964. Leaching
Raich, J. W.. E. B. Rastetter. J. M . Melillo. D. W. Kicklighter, of metabolites from foliage and subsequent reabsorption
P. A. Steudler. B. J . Peterson. A. L . Grace, B. Moore 111, and redistribution of the leachate in plants. American Jour-
and C. J. Vorosmarty. 1991. Potential net primary pro- nal of Botany 51:737-742.
ductivity in South America: application of a global model. Villela, D. M., and J. Proctor. 1999. Litterfall mass. chem-
Ecological Applications 1:399-429. istry, and nutrient retranslocation in a monodominant forest
Richter, D. D., and L. I. Babbar. 1991. Soil diversity in the on Maraca Island. Roraima, Brazil. Biotropica31: 198-21 1.
tropics. Advances in Ecological Research 21:315-389. Vogt, K. A , , C. C. Grier, C. E. Meier. and R . L . Edmonds.
Ruess, R. W., K . Van Cleve, J. Yarie, and L. A. Viereck. 1982. Mycorrhizal role in net primary production and nu-
1996. Contributions of tine root production and turnover trient cycling in Ahies amabilis ecosystems in western
to the carbon and nitrogen cycling in taiga forests of the Washington. Ecology 63:370-380.
Alaskan interior. Canadian Journal of Forest Research 26: Vogt, K . A , , D. J. Vogt, and J. Bloomtield. 1998. Analysis
1326-1336. of some direct and indirect methods for estimating root
Running, S . W., D. D. Baldocchi. D. P.Turner, S. T Gower, biomass and production of forests at an ecosystem level.
P. S. Bakwin, and K. A. Hibbard. 1999. A global terrestrial Plant and Soil 200:71-89.
Vogt, K . A., D. J. Vogt, P. A. Palmiotto, P. Boon, J. O'Hara.
monitoring network integrating tower fluxes, flask sam-
and H. Asbjornsen. 1996. Review of root dynamics in for-
pling, ecosystem modeling and EOS satellite data. Remote
est ecosystems grouped by climate, climatic forest type and
Sensing of the Environment 70: 108-127. species. Plant and Soil 187: 159-219.
Ryan, M. G . 1991. A simple method for estimating gross Waring. R. H.. and W. H. Schlesinger. 1985. Forest ecosys-
carbon budgets for vegetation in forest ecosystems. Tree tems: concepts and management. Academic Press, New
Physiology 9:255-266. York, New York, USA.
Ryan, M . G., R. M. Hubbard, S. Pongracic. R. J. Raison, and Weaver, P. L., and P. G. Murphy. 1990. Forest structure and
R. E. McMurtrie. 1996. Foliage, fine-root, woody-tissue productivity in Puerto Rico's Luquillo Mountains. Biotro-
and stand respiration in Pinus radiata in relation to nitrogen pica 22:69-82.
status. Tree Physiology 16:333-343. Yarie. J . 1997. Nitrogen productivity of Alaskan tree species
Santantonio, D., and J. C. Grace. 1987. Estimating fine-root at an individual tree and landscape level. Ecology 78:2351-
production and turnover from biomass and decomposition 2358.

View publication stats

You might also like