Вы находитесь на странице: 1из 20

Forest Ecology and Management 95 Ž1997.

209–228

A generalised model of forest productivity using simplified


concepts of radiation-use efficiency, carbon balance and
partitioning
a,) b
J.J. Landsberg , R.H. Waring
a
CSIRO Centre for EnÕironmental Mechanics, P.O. Box 821, Canberra, A.C.T. 2614, Australia
b
Oregon State UniÕersity, College of Forestry, CorÕallis, OR 7331, USA
Accepted 22 January 1997

Abstract

This paper describes a stand growth model, based on physiological processes, which incorporates a number of steps and
procedures that have allowed considerable simplification relative to extant process-based models. The model, called 3-PG
Žuse of Physiological Principles in Predicting Growth., calculates total carbon fixed Žgross primary production; PG . from
utilizable, absorbed photosynthetically active radiation Ž fp.a.u. ., obtained by correcting the photosynthetically active radiation
absorbed by the forest canopy Ž fp.a. . for the effects of soil drought, atmospheric vapour pressure deficits and stand age. PG
is obtained from fp.a.u. and the canopy quantum efficiency, values of which are becoming available. The ratio of net Ž PN . to
gross primary production is emerging as relatively constant for trees. This eliminates the need to calculate respiration and is
used to estimate P N —the net amount of carbon converted to biomass. 3-PG uses a simple relationship to estimate the
amount of carbon allocated below ground and a procedure based on allometric ratios—widely available for many species
and situations—to determine the allocation of carbon to foliage and stems and constrain tree growth patterns. The effects of
nutrition are incorporated through the carbon allocation procedure; the amount of carbon allocated below ground will
increase with decreasing soil fertility. Recently acquired knowledge about the physiological factors causing decline in forest
growth rates with age is used to model that decline. Changes in stem populations Žself-thinning. are derived from a
procedure based on the y3r2 power law, combined with stem growth rates.
The model requires weather data as input, works on monthly time steps and has been run for periods up to 120 years,
producing realistic patterns of stem growth and stem diameter increments. The time course of leaf area index is realistic for a
range of soil conditions and atmospheric constraints. 3-PG can be run from remotely-sensed estimates of leaf area index
coupled to weather data and basic, readily available information about soils and stand characteristics. It is being tested as a
practical tool against forestry data from New South Wales, Tasmania, Victoria and New Zealand. Test results show excellent
correspondence between stand growth measurements and simulated stem growth over 30 years. q 1997 Elsevier Science
B.V.

Keywords: Forest model; Carbon balance; Partitioning; Physiological processes; Weather

)
Corresponding author.

0378-1127r97r$17.00 q 1997 Elsevier Science B.V. All rights reserved.


PII S 0 3 7 8 - 1 1 2 7 Ž 9 7 . 0 0 0 2 6 - 1
210 J.J. Landsberg, R.H. Waringr Forest Ecology and Management 95 (1997) 209–228

1. Introduction called 3-PG Žthe acronym is derived from the use of


Physiological Principles in Predicting Growth. de-
A number of process-based models aimed at cal- veloped using this approach. The model is based on
culating forest productivity have been developed a number of well-established principles and some
within the last 10–15 years. Among these, FOREST- recently confirmed constants that greatly simplify
BGC ŽRunning and Coughlan, 1988; Running and calculations. It required little adjustment to obtain
Gower, 1991. is one of the best known and most realistic forest growth estimates from a simple
widely used, while others such as BIOMASS ŽMc- Žspreadsheet. program, and can easily be parame-
Murtrie et al., 1992., PnET ŽAber and Federer, 1992. terised for particular forest types. The constraints
and TREGROW ŽWeinstein et al., 1991. are also that lead to the actual amounts of intercepted photo-
well established. FOREST-BGC has been gener- synthetically-active radiation Žlight. that is utilised
alised to BIOME-BGC ŽRunning and Hunt, 1993. by forests are soundly-based in physiology and con-
and BIOMASS provided the basis for the G’DAY sistent with much more detailed models. This model,
model ŽComins and McMurtrie, 1993., which incor- moreover, can be applied to ground-based forest
porates nitrogen uptake and movement within trees inventory maps incorporated into a Geographical
and allows evaluation of the long-term consequences Information System ŽGIS., or to maps derived from
of changes in atmospheric CO 2 and the effects of the an analysis of satellite images, to extend estimates of
soil nitrogen balance on carbon accumulation by forest growth over large, heterogenous areas.
trees. All these models, however, are essentially The 3-PG model Žsee Table 1 for summary of
research tools and have yet to be simplified to the abbreviations and symbols.:
point where they are of value—or indeed of interest Ø calculates gross primary production Ž PG . from
—to practical forest managers. There is good reason utilisable absorbed photosynthetically active radi-
for this: the calculation of forest growth from physio- ation Ž fp.a.u. . and a canopy quantum efficiency
logical processes is complicated and has necessarily coefficient Ž a c .. fp.a.u. is obtained by reducing
involved the use of detailed, multi-variable models the values of absorbed photosynthetically active
that generally require a great deal of information and radiation Ž fp.a. . by amounts determined by modi-
careful parameterisation before they can be run. There fiers—dimensionless factors Ž f i . with values
is progress in simplification—FOREST-BGC has varying between zero and unity. The modifiers
been streamlined to the point where it requires only reflect the constraints imposed on the utilisation
an estimate of projected leaf area index Ž L) ., and of absorbed radiation by leaves because of sto-
appropriate weather data, making it suitable for use matal closure, caused by high atmospheric vapour
with satellite measurements ŽWhite and Running, pressure deficits Ž D ., soil drought, defined by the
1994., but the result remains an estimate of carbon ratio of the amount of water in the root zone to
fixed and biomass produced, which is of limited the maximum possible amount Ž u ., or the effects
value to those concerned with forest growth in the of sub-freezing temperatures ŽT . Žsee Landsberg,
conventional sense of changes in tree mass and the 1986; McMurtrie et al., 1994; Runyon et al.,
distribution of that mass to trees and their component 1994..
parts. Ø uses the ratio of net Ž P N . to gross primary pro-
We believe that the detailed models have now duction Ž PN rPG s c pp ., which is emerging as
developed to the point where they can be used to relatively constant for trees ŽSchulze et al., 1977;
generate simplified relationships, soundly based in Benecke and Evans, 1987; Ryan, 1991; Ryan et
biophysical processes yet accessible to practitioners al., 1996a; Williams et al., 1997; Heather Keith,
as well as to scientists. To generate practical tools personal communication, 1996. to estimate PN
these process-based calculations must be combined from PG .
with empirical relationships derived from experi- Ø uses a simple relationship, derived from informa-
ments and measurements made over long periods in tion in the literature about root growth and
forests and plantations. This paper describes a model turnover ŽSantantonio, 1989; Beets and White-
J.J. Landsberg, R.H. Waringr Forest Ecology and Management 95 (1997) 209–228 211

head, 1996. and the effects of growing conditions growth rates ŽYoder et al., 1994; Mencuccini and
on them ŽRunyon et al., 1994., to estimate the Grace, 1996; Ryan et al., 1996b. with age to
amount of carbon allocated below-ground. model that decline.
Ø uses a sub-model derived from the y3r2 power Section 2 presents an outline of the 3-PG model,
law Žsee Drew and Flewelling, 1977; Landsberg, which provides a framework for the description of
1986. and stem growth rates to calculate changes the simplifying concepts used in it. The ideas that we
in stem numbers per unit area with time Žself- believe have allowed us to make significant progress
thinning.. in calculating the carbon balance of, and carbon
Ø uses an equation based on allometric ratios— partitioning in forests, are dealt with in some detail
widely available for many species and situations in separate sections. Some results of sensitivity anal-
—to determine the allocation of carbon to foliage yses to illustrate the performance and potential of
and stems and constrain tree growth patterns. 3-PG are presented, as well as the results of compar-
Ø uses recently acquired knowledge about the isons between simulated and measured stem growth
physiological factors causing decline in forest rates.

Table 1
Abbreviations and symbols used in the text and in equations
e Radiation utilisation efficiency Žg MJy1 . based on above-ground standing biomass
ac Canopy quantum efficiency coefficient Žmol C Žmol photon.y1 .
hi Carbon allocation coefficients for foliage Žhf ., stems Žhs . and roots Žhr .
fs Short-wave incoming radiation ŽMJ my2 .
fp Photosynthetically active radiation Žmol my2 . Žf 0.5 fs .
fp.a. Absorbed photosynthetically active radiation ŽMJ my2 or mol my2 .
fp.a.u. Utilisable fp.a. ŽMJ my2 or mol my2 ., determined by environmental constraints
u Available water in the root zone Žmm, depth equivalent.
c pp Ratio PN rPG Ž0.45 " 0.05.
cg ,kg Empirical constant and coefficient in Eq. Ž4. describing litterfall during early growth
D Average monthly vapour pressure deficit ŽkPa.
fu Dimensionless modifier derived from average monthly soil moisture ratio
Fa Relative stand age Žactual agerestimated maximum age.
fage Dimensionless modifier based on stand age and associated decrease in stem hydraulic conductivity
fD Dimensionless modifier derived from average monthly vapour pressure deficit
fT Dimensionless temperature modifier based on number of frost days
gc Canopy conductance Žm sy1 .
g cm ax Maximum canopy conductance Žm sy1 .
PG Gross primary production Žcarbon fixed per unit time; mol my2 or Mg hay1 .
kg Empirical coefficient in Eq. Ž1., describing the relationship between stomatal and canopy conductance and D
ks Coefficient in the power-law equation describing change in maximum stem mass with stem population
L) Leaf area index
m Variable in Eq. Ž15., reflecting the effect of fertility on carbon allocation to roots
nu , cu Power and coefficient in Eq. Ž2., describing fu in terms of ru
n age Empirical power term in Eq. Ž3., for calculating the age modifier Ž fage .
PN Net primary production ŽMg hay1 . s Ž PG y autotrophic respiration.. In 3-PG, PN is assumed to be a constant
fraction Ž c pp s 0.45. of PG
pf.s Ratio of the growth rates of foliage and stem mass
ru Moisture ratio Žcurrent soil moisture contentru .
sf Specific leaf area Žm2 kgy1 ., based on projected leaf areas
W Total tree mass Žkg. at any time
wi Mass of component parts of trees: wf , foliage mass; ws , stem mass; wr , root mass
ws.max Maximum stem mass at a given stem population
212 J.J. Landsberg, R.H. Waringr Forest Ecology and Management 95 (1997) 209–228

2. Model structure ing values can be obtained from the literature, or


reasonable guessed values will suffice. L) is deter-
A flow diagram outlining the calculations that mined from foliage mass and input values of specific
comprise the model is presented in Fig. 1. The model leaf area Ž sf , m2 kgy1 ., which are widely available.
uses a monthly time step and requires as input data fp is assumed to be 0.5 fs and the model calculates
Ø values of total short-wave Ž375–2500 nm. incom- fp.a. using Beer’s law. This assumes that foliage is
ing radiation Ž fs .; uniformly distributed across the stand, which is not
Ø monthly mean day-time vapour pressure deficit an unreasonable assumption from the time stands are
Ž D .; approaching canopy closure Žfor comparisons be-
Ø total monthly precipitation; tween the performance of detailed and simple radia-
Ø number of days per month with frost. tion interception models, integrated over time see
Starting values of foliage, stem and root mass are Wang et al., 1992.. It is a poor assumption for
also required, appropriate to the age of the stand at widely-spaced young trees, so radiation interception,
the start of a run, together with appropriate allom- and hence rates of dry mass production, calculated
etric equations and some soil water parameters. Start- by 3-PG for young forests, may be significantly in
error. Whether this is important over the life of a
stand requires investigation; the matter can be rela-
tively easily rectified by using a more complex
radiation interception model, at least for early stage
growth Žsee, for example the procedure used by
Kuuluvainen, 1991..

2.1. Vapour pressure deficit modifier ( f D )

If L) - 3, canopy conductance Ž g c . is calculated


as maximum stomatal conductance Ž g cmax . Žfor val-
¨
ues see Korner, 1993. corrected for the effects of D
Žspecified by f D . multiplied by L) . If L) ) 3, canopy
conductance is derived from maximum conductance
values ŽKelliher et al., 1993, 1995. corrected for the
effects of D. The correction applied to g c is given
by
g c s g cmax exp Ž yk g D . Ž 1.
where the coefficient k g is based on well-established
relationships between stomatal conductance and
vapour pressure deficit Žsee, e.g. Dye and Olbrich,
1993; Leuning, 1995; Granier et al., 1996.. We used
a value of k g s 2.5 Žwith D in kPa.. The model is
sensitive to this relationship. Eq. Ž1. is also used to
calculate the D modifier applied to fp.a., i.e. f D s
expŽyk g D ., so that f D ™ 1 as D ™ 0. Support for
the applicability of this relationship directly to fp.a.
has been provided by Landsberg and Hingston
Fig. 1. Flow diagram showing the sequence of calculations in Ž1996., who showed that monthly growth increments
3-PG. The model was developed using a spreadsheet and ‘Visual
Basic’ and exists in this form. It has also been programmed using
of plantation eucalypts were negatively correlated
a powerful modelling package, which runs much faster and can be with monthly average D, and that the relationship
used in association with satellite data and GIS. between growth and fp.a. was improved by correct-
J.J. Landsberg, R.H. Waringr Forest Ecology and Management 95 (1997) 209–228 213

ing for the effects of D. They obtained a value of


k g s 1.9.

2.2. Soil water modifier ( fu )

The soil water balance is obtained as the differ-


ence between total monthly transpiration Žmm., cal-
culated using the Penman–Monteith equation with
the appropriate g c value, and monthly precipitation.
The model is initialised with soil water contents
maximum available waters Ž u mm. in the rooting
zone. This is dependent on the water holding charac- Fig. 2. Relationship between the soil water modifier Ž fu . and the
teristics of the soil and the rooting depth of the trees moisture ratio Ž ru . for four soil types.
Žsee Landsberg and Gower, 1997. The moisture ratio
Ž ru . for the stand is calculated as
Landsberg and Gower Ž1997. for a detailed discus-
Current soil water contentq water balance sion.. The values of cu and nu are, nevertheless,
ru s chosen without specific empirical justification. If
Available water
appropriate experimental results become available
The water balance in any month will be negative if these values may need to be altered.
transpiration exceeds precipitation, and vice versa. If
the numerator of the expression for ru exceeds u , it 2.3. Temperature (frost) modifier ( f T )
is set to u , i.e. the excess water is assumed to have
run off or drained out of the system. If it is negative, The present version of the model accounts for
ru s 0. Available water can be set to any value temperature only in terms of the occurrence of frost.
considered appropriate to the forest under considera- We assume that there is no photosynthesis on any
tion. Values in the range 25–250 mm occur in the day that temperatures fall below zero Žthis could
literature ŽWaring and Major, 1964.. easily be altered, on the basis of appropriate empiri-
The soil water modifier, fu , is calculated from cal evidence, to more than 1 day.. The frost modifier
1 is, therefore
fu s nu Ž 2.
1 q Ž 1 y ru . rcu f T s 1 y Ž frost days per month

where cu and the power nu take different values for rnumber of days per month.
different soil types. We suggest cu s 0.7, 0.6, 0.5
and 0.4 for sand, sandy-loam, clay-loam and clay, 2.4. Nutrition
respectively, and nu s 9, 7, 5 and 3 for the same soil
types. These values produce the curves shown in Fig. A number of models have used the relationship
2, which are essentially the same as those for relative between leaf photosynthesis and nitrogen concentra-
transpiration rate vs. volumetric water content pub- tion to scale carbon assimilation at stand levels.
lished by Denmead and Shaw Ž1961. and, much However, this relationship is not well-founded for
later, by Dunin et al. Ž1985. for Eucalyptus macu- conifers Žsee Landsberg and Gower, 1997. and we
lata. The curve shift from sand to clay, causing fu to have chosen to deal with nutrition through the mech-
fall earlier in clay, but not as rapidly as in the sand, anism of carbon allocation rather than attempt to
reflects the differences in the hydraulic character- calculate a nutritional modifier that would require
istics of soils with different proportions of clay. At a information on nutrient availability, and uptake by
given soil water content the water potential of clay trees, which would be extremely difficult to simu-
soils is significantly lower than in sandier soils Žsee late. Detailed discussion of this is deferred until after
Williams et al. Ž1983. for data illustrating this, and the section on carbon allocation.
214 J.J. Landsberg, R.H. Waringr Forest Ecology and Management 95 (1997) 209–228

2.5. The age effect ( f a g e .

As a forest ages, its above-ground net primary


production decreases, a fact well documented in all
forestry yield tables. Until recently, the observed
decrease in PN with age was thought to be related to
an increase in maintenance respiration of woody
biomass, but direct measurements of stem respiration
at times when growth is not occurring suggest that
less than 12% of annual PG is required for mainte-
nance of the small number of living cells present in
woody tissue, even in subtropical climates ŽRyan et
al., 1995.. Some have speculated that above-ground
Fig. 3. The linear relationship between hydraulic conductance and
production decreases because soil nutrient supply P N obtainable from the data published by Mencuccini and Grace
declines as more woody material is incorporated into Ž1996.. The low values of P N and conductance were measured in
litter and soil organic matter and nutrients are in- old stands and the highest values in young stands. The relationship
creasingly immobilised ŽGower et al., 1996; Murty et supports the assumption that stem conductance declines with age,
al., 1996.. However, nutrient availability in old stands inducing lower stomatal conductance and hence lower photo-
synthesis and stand productivity.
can be similar to or greater than in younger stands
ŽRyan et al., 1996b.. Paired young and old-tree
comparisons made on the same site Žwith, presum-
ably, similar nutrient availability. showed that older sion, in terms of relative stand age, to account
trees had similar maximum rates of photosynthesis in explicitly for the reduction in maximum stomatal
the early morning but during the day these rates fell conductance as stands age. The age modifier Ž fage .,
25–30% more than those observed on the foliage of is given by an equation that mimics the changes in
younger trees. The differences were attributed to Žnormalised. above-ground wood production shown
differences in the relative sensitivity of stomata to by data collated by Ryan et al. Ž1996b.. Relative age
atmospheric vapour pressure deficits ŽYoder et al., Ždenoted Fa . is the ratio of actual age Žin years. to
1994., and appear to be associated with hydraulic the maximum age likely to be attained by a forest or
limitations to the flow of water through an increas- plantation
ingly long and torturous path as trees age ŽTyree and 1
Sperry, 1988; Ryan and Yoder, 1997.. This interpre- fage s n age Ž 3.
1 q Ž Far0.95 .
tation is supported by data presented by Mencuccini
and Grace Ž1996. who showed that, in Scots pine The constant Ž0.95. causes fage s 0.5 when Fa s
stands differing more than a half-century in age, 0.95; the power in the denominator controls the rate
changes in the total hydraulic conductance in stems of change of the function. We have used n age s 4.
and branches closely parallelled annual variation in Either of these parameter values can be varied as
above-ground net primary productivity as the forest deemed appropriate by users. The expression is not
aged ŽFig. 3.. Hydraulic conductance, calculated on sensitive to the maximum age specified for a forest,
a unit leaf area basis, was lower by a factor of more although it will clearly cause large differences in
than four in mature trees compared to the values in growth rates of stands that are, say, 20 years old, if
stands at maximum productivity. Because of the the maximum age is specified as 80 years—as may
reduction in stomatal conductance associated with be the case for some plantations in sub-tropical areas
reductions in hydraulic conductivity as trees age, —or 300 years, which would be a normal age for a
photosynthesis is reduced, and less carbon is avail- mature forest in many parts of the world.
able to maintain the previously acquired leaf area. Regardless of growing conditions, Eq. Ž3. results
The canopy, as a result, must open as a forest ages. in patterns of biomass production that decrease at
In the 3-PG model we use an empirical expres- rates comparable to the reported decrease in above-
J.J. Landsberg, R.H. Waringr Forest Ecology and Management 95 (1997) 209–228 215

ground production published in forestry yield tables portion to the break-down in chlorophyll pigmenta-
appropriate for a given region. tion, a change easily assessed visually from the
ground or from satellites ŽYoder and Waring, 1994;
2.6. Calculation of f p. a.u. Prince and Goward, 1995.. In another study, Williams
et al. Ž1997. computed PG for predominantly ever-
Utilizable, absorbed photosynthetically active ra- green forests distributed across western Oregon by
diation is calculated by applying the modifiers for D adding estimates of foliage, stem, and root respira-
Ž f D ., soil water Ž fu ., temperature Ž f T . and age Ž fage . tion to previously published estimates of PN and
to fp.a.. The soil water and vapour pressure deficit fp.a. ŽRunyon et al., 1994.. They confirmed that the
modifiers are not multiplicative relative to each other. maximum canopy quantum efficiency has a value for
3-PG uses the lower value of f D and fu , on the forests near 0.03 mol C Žmol photon.y1 and does not
assumption that if stomata are closed by D, limiting vary widely about that value. On severely nutrient
transpiration rate Žand CO 2 uptake. so that the rate deficient soils, or where atmospheric pollutants cause
of water movement to the roots is fast enough to chlorosis of foliage, a c will decrease below the
maintain the transpiration rate even in relatively dry maximum ŽJohnson and Linberg, 1992.. Tests of the
soil, then soil water content is not a limiting factor. constancy of a c can be expected from any forest
The reverse holds: if water cannot move to roots fast experiment involving long-term CO 2 flux measure-
enough to meet transpiration demand then g c must ments, or from experiments where most of the terms
be reduced to a value appropriate to the supply rate, in the dry mass balance are accounted for and a
even if that is below that determined by D, otherwise reliable process-based model is used to calculate PG .
the trees would become desiccated. In this case soil The use of a linear relationship between fp.a. and
water is the limiting factor, not canopy conductance. dry mass production by plant canopies, stemming
largely from Monteith Ž1977., is now widespread
2.7. Dry mass production (radiation conÕersion effi- and there are many values of the dry mass conver-
ciency) sion factor Ž e , g MJy1 or mol C Žmol photon.y1 . in
the literature Žsee review by Landsberg et al., 1996..
PG is calculated by multiplying fp.a.u. by the It is often not entirely clear, however, whether the
canopy quantum efficiency coefficient a c Žsee be- analyses refer to above-ground or total biomass pro-
low., and PN follows. Carbon allocation is calcu- duction in relation to fp.a.. When above-ground
lated monthly Žsee Section 2.11. and L) for 1 month biomass is referenced there is always uncertainty
is derived from the leaf mass at the end of the about the below-ground component. Large variations
previous month. in e can be explained on this basis alone. Further-
3-PG uses a ‘universal’ canopy quantum effi- more, many analyses leading to values of e have not
ciency coefficient a c , with a value of 0.03 mol C taken account of constraints on radiation utilisation
Žmol photon.y1 , equivalent to 1.8 g C MJy1. This is by plant canopies. 3-PG allows analysis of the above
based on three studies. McMurtrie et al. Ž1994. and below-ground biomass fractions, so that e val-
analysed pine productivity at five sites ranging from ues can be examined. We propose that the symbol e
Sweden through the USA to Australia and New be used only for above-ground P N per unit utilisable
Zealand, and found simulations of PG at those sites fp.a. and the symbol a c be used for PG .
produced values which fell on a straight line with a The relationships are conceptually simple. Stand-
slope of 1.77 g C MJy1 when plotted against fp.a.u.. ing biomass at any time is W Ž t . —the integral of the
Waring et al. Ž1995., in conjunction with an indepen- processes of dry mass production Ž PG . and respira-
dent estimate of gross primary production derived tion Ž R . up to that time, i.e.
from eddy correlation data collected over the Har-
W Ž t . s PG y R Ž 4.
vard Forest, a mixed deciduous forest in north-east
USA, showed the average value of maximum canopy so that
quantum efficiency to be 0.03 mol C Žmol photon.y1 .
Reductions in a c occurred seasonally in direct pro- a c s PG rfp .a .u . Ž 5.
216 J.J. Landsberg, R.H. Waringr Forest Ecology and Management 95 (1997) 209–228

and k s is a coefficient for which a value is required. This


e s Ž 1 y pa . PG rfp .a .u . Ž 6. can be obtained either from empirical data—the
average mass of individual stems in stands near the
where pa is the mass fraction Žintegrated over time. end of their life cycle—or a value can be obtained
allocated to root growth and turnover. from 3-PG by running the model through an appro-
priate rotation length with low stocking Ž100–150
2.8. Litterfall trees hay1 . with the mortality function ‘switched
off’, to estimate maximum achievable stem mass for
It is important to include litterfall in the model. the environmental conditions and allometric ratios
Any value considered appropriate can be used; we used in the simulation.
find that a constant monthly litterfall rate Žg f.max . of To run the population sub-model, initial stem
about 0.02 Ži.e. about 25% per year. gives good mass is provided as an input, and from the end of
results. It would be simple to make litterfall a vari- Year 1 mean stem mass Ž ws . generated by stem
able function of some cumulative drought index Žsee, growth in the model is tested against ws,max for the
e.g. experimental results of Pook, 1986; Linder et al., current population Ž pŽ ws.max ... If ws ) ws.max we
1987.. We have not done this but we have made solve Eq. Ž8. for pŽ ws . and reduce stem numbers by
litterfall a function of age for young stands—increas- D p s w pŽ ws . y pŽ ws.max .x, i.e. D p is stem mortality.
ing asymptotically to the maximum Ž‘standard’. lit- The procedure is illustrated diagrammatically in Fig.
terfall rate from near zero at 1 year to the maximum 4. By the end of the following year, mean stem mass
at 5 years ŽEq. Ž7.. will have increased and the procedure is repeated. It
g f .max has the disadvantage that stem numbers fluctuate
gf s Ž 7.
1 q cg exp Ž ykg t . above the ws.maxrp line, instead of below it Žsee
Drew and Flewelling, 1977; Landsberg, 1986., but in
where g f is the foliage Žand small branch. litterfall
view of the approximations involved in the solution
rate Žyeary1 ., cg is an empirical constant, kg is an
for k s , this is not important.
empirical coefficient and t is time, in months. We
set cg s 15 and kg s 0.12.
3-PG includes root turnover rates, although these
have no effect on its general performance because
we do not use root mass to modify estimates of
rooting depth. It would be a simple procedure to do
this if it were considered justified by the knowledge
available about a particular system.

2.9. Stem populations

The model requires stem numbers because the


allometric ratios used to partition carbon are invari-
ably determined for single trees Žsee Section 2.11.;
also, forest managers are not interested in the total Fig. 4. Diagrammatic representation of the stem population sub-
biomass produced by a forest—they want to know model. The line describes the relationship between average maxi-
stem growth rates and final stem volumes. 3-PG mum individual stem mass Ž ws.max . and population. If mean stem
exploits the Žvirtually universal. relationship between mass Ž ws . is less than ws.max Žpoint 1. the population remains
stem populations and maximum achievable individ- unchanged. If ws ) ws.max Žpoint 2., p is reduced by the number
necessary to conform to the relationship. Values of k s have to be
ual stem mass Ž ws.max ŽEq. Ž8..; see Drew and provided. They can be obtained from empirical data or calculated
Flewelling, 1977; White, 1981; Landsberg, 1986. to using the model. We calculated k s from stem mass values ob-
calculate changes in stem populations Ž p . tained by running the model for 120 years using initial stem
populations of 150 hay1 , with the normal environmental con-
ws .max k s py3 r2 Ž 8. straints.
J.J. Landsberg, R.H. Waringr Forest Ecology and Management 95 (1997) 209–228 217

The mass of the stems that die must be discarded ‘harshness’ of the environment: under zero-stress
from standing biomass. Stem mortality largely oc- conditions f D , fu and the temperature modifier Ž f T .
curs among the smaller stems in a population so in are unity, and fp.a.u.s fp.a., so operationally we
this version of the model we divide D p by 3 and define the harshness of the environment in terms of
calculate stem mass loss as Ž D pr3. wsŽ1r3 q 1r6 q the ratio of fp.a.u. to fp.a.—i.e. the smaller the
1r9.. We do not reduce leaf mass, on the assump- proportion of radiation absorbed by foliage that could
tion that dying stems would have few leaves, and be utilised, the more harsh the environment ŽEq.
that the stand foliage mass is distributed among the Ž11.; Section 2.11..
living stems.
2.11. Carbon allocation
2.10. Respiration, PN and root turnoÕer
Carbon products derived from photosynthesis in
Ryan’s work, in recent years ŽRyan, 1991; Ryan leaves are transported to the various parts of trees,
et al., 1995, 1996a., has contributed greatly to our where they are used for respiration and as structural
knowledge of the respiration rates of trees and their material. The mechanisms determining the amount of
various component parts. It has resulted in useful carbon allocated to any part of a tree are not well
relationships between respiration rates and variables understood; study of these processes remains one of
such as temperature and leaf nitrogen content but, the major challenges facing physiological ecologists
because they require information such as sapwood and various models have been developed to describe
volume, fine root mass, tissue temperatures and ni- ˚
them Žsee Agren, ˚
1983; Agren and Ingestad, 1987;
trogen content, these are of limited value in stand- Thornley, 1972a,b and the review by Cannell and
level models. However, by combining these studies Dewar, 1994..
with increasingly reliable estimates of PG , derived In models aimed at simulating the growth of real
from measurements ŽSchulze et al., 1977; Benecke forests, as opposed to exploring the mechanisms that
and Evans, 1987; Waring et al., 1995; Williams et influence growth patterns, carbon allocation is some-
al., 1997. and models such as BIOMASS ŽRyan et times estimated using allocation coefficients Žhi . de-
al., 1996a. new information about the relationship rived from the Žallometric. equations that describe
between PG and PN is emerging, from which it the observed relationships between the mass or size
appears that the ratio PN rPG Žs c pp . is remarkably of different parts of plants Žsee, e.g. McMurtrie and
constant, averaging about 0.45 " 0.05 for a wide Landsberg, 1992.. These allocation coefficients are
variety of forests, including deciduous hardwoods dimensionless, but the problem of their dynamics has
and evergreen conifers, both young and old. ŽThis not, up to now, been solved.
value was derived from data in the papers cited Allometric relationships for single trees can, al-
above, and from data made available to us by Michael most invariably, be described by equations of the
Ryan, personal communication, and—relating to eu- form
c a ly p ts — b y H e a th e r K e ith , p e rs o n a l
Wi s a iW n i Ž 9.
communication..
Two important generalisations have emerged from where W is the total mass of the plant and i denotes
recent analyses of carbon allocation in trees: Ž1. any component part. It is reasonable to assume that
there is a strong inverse relationship between stem the parameter values—particularly n i —reflect the
growth and the fraction of PN allocated below ground genetic characteristics of species. There are many
ŽBeets and Whitehead, 1996.; Ž2. as environmental examples in the literature where experimentalists
conditions become harsher, the fraction of P N allo- have obtained improved descriptions of data by fit-
cated annually to fine root growth increases from ting equations of other forms, usually with more
about 25% to nearly 60% ŽSantantonio, 1989; Run- parameters, but the fact remains that almost all data
yon et al., 1994; Beets and Whitehead, 1996; Heather sets can be described, with high statistical r 2 values,
Keith, personal communication.. In 3-PG the envi- by Eq. Ž9., which is a very tractable and useful form
ronmental modifiers provide a measure of the Žsee Pearson et al., 1984; where a summary of data
218 J.J. Landsberg, R.H. Waringr Forest Ecology and Management 95 (1997) 209–228

for Pinus contorta indicates remarkable stability Eqs. Ž10. – Ž12. were originally written in that
across stands with different stem populations.. More form by McMurtrie and Wolf Ž1983.. They have
complex statistical descriptions seldom contribute since been used in numerous models; see, e.g. Mc-
much additional useful information. Foresters nor- Murtrie Ž1985. and McMurtrie and Landsberg Ž1992.
mally use stem diameter at ‘breast height’ Ž B, for- for an application to analysis of experimental data.
mally B f 1.4 m. as a measure of stem size; this can To obtain the allocation coefficients for 3-PG we
be used as a surrogate for W. The values of the specified the way hr is expected to vary with grow-
parameters of Eq. Ž9. may be expected to vary with ing conditions Žsee Section 2.10.. The equation used
age Žsee Table 2. and site water and fertility condi- is a simple hyperbola
tions, although there are few studies available where
these factors have been examined. However, the 0.8
hr s Ž 13 .
empirical values available for the power Ž n i . are 1 q 2.5 Ž fp .a .u . rfp .a . .
remarkably similar Žfor a summary of experimental
data from many sources see Gholz et al., 1979; making hr dependent on the relative harshness of the
Pastor et al., 1984 and Landsberg, 1986 for discus- growing conditions, defined by the ratio fp.a.urfp.a .
sion.. Eq. Ž13. reflects the accumulating evidence that trees
If the net rate of dry mass production of a tree is allocate increased amounts of carbon to their roots
dWrdt, then the rates of growth in leaf Ž wf ., root when growing conditions deteriorate—a strategy that
Ž wr . and stem Ž ws . mass are seems intuitively ‘sensible’ in terms of the need to
dwf dW survive adverse conditions—and imposes limits to
sh y g f wf Ž 10 . the proportion of carbon allocated to root growth,
dt dt f maintenance and turnover that are consistent with
dwr dW many studies and observations and physiologically
h y gr wr Ž 11 . plausible.
dt dt r
dws dW It is clear from Eqs. Ž10. – Ž12. that the allocation
s h y gs ws Ž 12 . coefficients are, essentially, given by the ratioŽs.
dt dt s Ždwirdt .rŽdWrdt . s dwrdW so, accepting B as a
where g f , gr and gs represent litterfall, root turnover good surrogate for W, we can assume that the alloca-
and stem mortality rates Žcf. Eq. Ž7. and the outline tion coefficients for foliage and stems are propor-
of the procedure for calculating stem mortality, fol- tional to dwfrd B and dwsrd B. The ratio of the stem
lowing Eq. Ž8... and leaf allocation coefficients Žapplicable to PN .

Table 2
The parameters of Eq. Ž9. Ž wi s a i Bin . describing the relationships between the mass of foliage Ž wf , kg. and stems Ž ws , kg; this includes
branches and bark. and diameter at breast height Ž B, cm. for a number of species and—in the case of eucalypts—different age classes. The
data are intended to be illustrative, not comprehensive
Species af nf as ns Data source
E. globulus 2 years 0.252 1.11 0.257 1.66 Robin Cromer, CSIRO
E. globulus 4 years 0.036 1.98 0.173 2.02 Robin Cromer, CSIRO
E. globulus 6 years 0.013 2.45 0.126 2.23 Robin Cromer, CSIRO
E. globulus 9 years 0.010 2.42 0.091 2.48 Robin Cromer, CSIRO
E. pauciflora 55 years 0.050 1.61 0.031 2.61 Dr Heather Keith, CSIRO
Pinus spp. Various ages 0.009 2.32 0.040 2.65 Gower et al. Ž1994. a
Pinus contorta Various popns. 0.018 2.06 0.153 2.27 Pearson et al. Ž1984.
Pinus contorta ? 0.027 1.84 0.050 2.43 Gholz et al. Ž1979.
Acer saccharum About 60 years 0.006 2.22 0.073 2.56 Pastor and Bockheim Ž1981.
Populus tremuloides About 60 years 0.024 1.50 0.171 2.20 Pastor and Bockheim Ž1981.
a
Calculated from data presented by Gower et al., including various pine species of different ages grown in various locations
J.J. Landsberg, R.H. Waringr Forest Ecology and Management 95 (1997) 209–228 219

can, therefore, be obtained from Eq. Ž9. with numeri-


cal values of a i and n i appropriate to each compo-
nent of the treeŽs. and the species under considera-
tion. This gives us
Ž dwfrd B .
pf .s s Ž 14 .
Ž dwsrd B .
where pf.s is the ratio of the growth rates of foliage
and stems, in terms of their changes in relation to B.
Now, given that hf q hr q hs s 1 and that we have
hr from Eq. Ž13., then
Fig. 6. Variation of the carbon allocation coefficient Žsee Eq. Ž13..
hs s Ž 1 y hr . r Ž pf .s q 1 . and hf s 1 y hr y hs with site nutritional status, or nutrient availability, introduced as a
modifying variable. The curves are described by Eq. Ž15..
Operationally, in 3-PG, average stem mass is calcu-
lated from total Žstand. stem mass and stem number.
Eq. Ž9. is inverted and solved for B to provide the
value for use in Eq. Ž14.. Foliage and stem mass are to the fine roots, but a minimum of about 20% goes
never calculated directly from Eq. Ž9.. to coarse roots ŽBeets and Whitehead, 1996.. In
This procedure has proved remarkably stable and general, climatic constraints over-ride nutritional lim-
well-behaved. If PN is reduced by poor growing itations—trees in the most fertile soil will not grow
conditions, hr is increased ŽEq. Ž13.. and stem and without adequate water, or in very cold conditions
foliage growth are reduced in a way that conserves Žsee Linder et al., 1987; Specht and Specht, 1989;
the allometric balance of the trees. Snowdon and Benson, 1992. for examples of data
illustrating the way water relations can over-ride
2.12. Effects of nutrition nutrition. —but, when water is adequate, improved
nutrition usually results in increasing the proportion
There is evidence that nutrition affects the amount of P N allocated above ground.
of carbon allocated to roots Žsee review by Santanto- Quantifying the relationship is difficult—our un-
nio, 1989; Fig. 5., with a greater proportion going to derstanding of nutrient dynamics and uptake pro-
the roots of trees on infertile sites than those on cesses is not yet good enough, at the treerstand
fertile sites. Of the carbon going to roots, most goes level, to allow formulation of definitive, determinis-
tic relationships so we have used a simple scaling
procedure, whereby the allocation of carbon below
ground is minimised on fertile sites and increases as
fertility decreases. The proposed relationships are
presented in Fig. 6, which presents a set of curves
derived from Eq. Ž13. with a new variable Ž m.
introduced:
0.8
hr s Ž 15 .
1 q 2.5m Ž fp .a .u . rfp .a . .
m has a maximum value of 1, on a highly fertile site,
reducing towards zero as site fertility decreases. The
Fig. 5. Relation between total nitrogen content of Pinus radiata curves in Fig. 6 indicate that differences in carbon
canopies and the fraction of P N allocated to roots. The data were allocation caused by soil fertility are greatest under
obtained from experiments on 7–11 year old trees reported by
Beets and Madgwick Ž1988. and Beets and Whitehead Ž1996..
good growing conditions Ž fp.a.u. rfp.a.™ 1. and least
The line was drawn by eye to indicate that these data are when growth is restricted by vapour pressure deficits,
consistent with the relationships described by Eqs. Ž13. and Ž14.. soil water or temperature. We can only suggest, at
220 J.J. Landsberg, R.H. Waringr Forest Ecology and Management 95 (1997) 209–228

Fig. 7. Climatic data from Canberra, Australia, used in all the simulations reported in this paper. The lightly shaded columns are monthly
rainfall Žmm., the more intensely shaded columns are monthly incoming radiation ŽMJ my2 dayy1 . and the solid columns are average
vapour pressure deficit Ž D ., here given in millibars for presentational reasons Ž10 mbar s 1 kPa.. The hatched area indicates frost days per
month.

this stage, that values of m be allocated on the basis biomass values were 1, 3 and 6 t hay1 dry mass for
of measures of fertility such as litterfall nitrogen or foliage, roots and stems, respectively. The soil was a
nitrogen mineralisation rates, as well as conventional sandy-loam Žsee Eq. Ž2. and Fig. 2.. All runs were
analyses of soil phosphorus content Žsee Landsberg carried through to 100 years.
and Hingston, 1996.. Fig. 8 shows the leaf area index changes Ž L)

3. Results

We present, in this section, results of sensitivity


analyses with 3-PG and some results from tests of
model output against measured stand growth.
The model may be run with observed climatic
data Žmonthly values for a number of years. or run
for as many years as required using the same set of
data. For illustrative purposes in the sensitivity anal-
ysis we have chosen the latter option, using climatic
data from Canberra, in the Australian Capital Terri-
tory, shown in Fig. 7. This allows us to explore the
behaviour of the model without confounding the
Fig. 8. Trends in leaf area index Ž L) . under different environmen-
comparisons by using different weather data for each tal constraints. All simulations were performed using the Canberra
year. We have not varied nutrition; the stands are climatic data, with initial stem populations Žstocking density. of
assumed to be in fertile soils Ži.e. m s 1 in Eq. Ž14.., 1000 hay1 . The environmental conditions relating to each curve
but m was varied to account for soil fertility to give are given in Table 3. Note that frost has relatively little effect
the results of the model tests presented in Fig. 11. because radiation during winter months is low. Increasing soil
water storage has some effect, but limitations imposed by vapour
In all the simulations shown here the allometric pressure deficits still dominate during months with high radiation.
ratios used were for generic pine Žsee Table 2., but Removing the vapour pressure deficit limitation results in more
with n f s 2.42, unless otherwise stated. Starting rapid exhaustion of the soil water supply.
J.J. Landsberg, R.H. Waringr Forest Ecology and Management 95 (1997) 209–228 221

Table 3
Stem growth data relating to Fig. 8, where trends in L) under different environmental conditions are shown. The numbers in the left-hand
column correspond to the numbers on the curves in the figure
Run Constraints Max. growth rate Max. L) Final stem biomass Final stem number
ŽMg hay1 yeary1 . ŽMg hay1 . Žno. hay1 .
1 None 31.2 9 2217 226
2 D; u non-limiting, frost 10.8 3.8 773 199
3 D; u non-limiting, no frost 9.7 3.1 728 239
4 No D; u s 150 mm, frost 8.1 2.7 680 345
5 D, u s 150 mm, frost 8.2 2.7 601 229

values at the end of the last month of each year. with always unconstrained, but not very different in the
time. The top line ŽCurve 1. shows the time course other cases.
of L) when growth was not constrained either by In Fig. 9 we explore the consequences of chang-
soil water or D; the environmental conditions relat- ing the allocation of dry mass to leaves by changing
ing to the other curves are given in Table 3. The the exponent for foliage Žhf in Eq. Ž9.. from 2.42,
main point that emerges from Fig. 8 is the very
strong constraint imposed on the development of leaf
area by water limitations, whether caused by short-
age of soil water or by high atmospheric vapour
pressure deficits. The stem growth data from these
runs are given in Table 3, where we see that maxi-
mum growth rates, which occur early, reflect u
values before stored soil water Žset to u at the start.
is exhausted and the water balance is dominated by
rainfall. However, high growth rates in the first
10–15 years do not necessarily result in high final
yields. These were very high where growth was

Fig. 10. Ža. Changes in stem population with different initial stem
numbers: 1, 5000 stems hay1 ; 2, 2500 stems hay1 ; 3, 1000 stems
hay1 . Note that stem populations begin to fall at different times.
Žb. Net stemwood growth rates at the three populations used in
Ža.. Note that, at the two high populations Žcurves 1 and 2, starting
Fig. 9. Stem growth rates calculated with different values of n f values 5000 and 2500 stems hay1 , respectively., stemwood growth
ŽEq. Ž9.., which determines allocation of carbon to foliage growth. rate begins to fall after about 5 years, falling faster in 1. The low
The values used and environmental conditions relating to each population stand Žstarting value 1000 stems hay1 . sustained high
curve are given in Table 4. growth rates for much longer, and ended with higher yields.
222 J.J. Landsberg, R.H. Waringr Forest Ecology and Management 95 (1997) 209–228

Table 4
Stem growth data relating to Fig. 9, where the stem growth patterns resulting from the use of different allometric ratios, determined by n f
ŽEq. Ž8.., with and without environmental constraints, are illustrated. The numbers in the left-hand column correspond to the numbers on the
curves in the figure
Run Constraints nf Max. growth rate Max L) Final stem biomass Final stem number
ŽMg hay1 yeary1 . ŽMg hay1 . Žno. hay1 .
1 None 2.42 32.8 9 2217 226
2 None 2.32 30.0 4.7 1960 308
3 D, u s 150 mm, frost 2.42 8.2 2.7 601 229
4 D, u s 150 mm, frost 2.32 8.0 1.5 320 127

used in all other runs, to 2.32. The exponent for tions over 100 years with u s 150 mm and vapour
stems Ž n s . remained at 2.65 Žsee Section 2.11.. The pressure deficit constraints. Fig. 10Žb. presents the
upper lines show that, in the absence of environmen- respective stem growth rates Žnet of stem losses.
tal constraints, allocating a larger proportion of avail- over 100 years with different initial stocking levels
able carbon to leaves Ž n f s 2.42. results in slightly Ž1000, 2500 and 5000 stems hay1 .. The population
higher growth rates and final yield ŽTable 4. than the curves follow the course determined by Eq. Ž8., from
lower allocation rate Ž n f s 2.32.. Stem mortality, the time ws s ws.max . Stemwood growth rates were
which starts when average stem weights reach ws.max , sustained at relatively high levels for about 35 years
begins about the same time. The lower lines Ž3 and at the lowest population Ž1000 hay1 ., but fell quite
4. were calculated with u s 150 mm and normal early at the higher populations.
vapour pressure deficit constraints. Here we find that Fig. 11 shows comparisons between 3-PG output
the greater allocation of carbon to leaves is an and the growth of P. radiata at Tumut, in New
advantage. Maximum growth rates are similar but South Wales, and Haupapa, in North Island, New
higher growth rates are sustained for longer by the Zealand over 30 years. Tumut Ž35.258S, 148.58E. is
canopy with the greater leaf area and the final yield at an elevation of 1100 m, with average precipitation
is almost double that achieved with n f s 2.32. The of 1360 mm yeary1 and incoming solar radiation
biomass data for these curves are given in Table. 4. Ž fs . of 7540 MJ my2 . Frost occurs on about 120
Fig. 10Ža. shows the reduction in stem popula- days yeary1 . The soils at Tumut are poor and were
rated 0.3 on a scale of 0–1 Ži.e. in Eq. Ž15., m s 0.3.,
with an average Žestimated. 150 mm of water avail-
able in the root zone. Haupapa Ž38.58S, 176.48E. is
at an elevation of 580 m, with average precipitation
of 1150 mm yeary1 and fs s 7615 MJ my2 yeary1 .
There are less than 10 frost days per year. The soils
are excellent, rated m s 1, with 400 mm water avail-
able in the root zone. For the simulations initial
stocking was set to 1500 hay1 for both sites—the
normal commercial planting density. The k s value
ŽEq. Ž8.. which determined natural mortality, was set
to 6 = 10 6 at both sites, although it should perhaps
Fig. 11. Simulated Žlines. vs. measured values Žpoints. of cumula- have been higher for Haupapa. The data with which
tive stem biomass at Haupapa, New Zealand, and Tumut, New simulated values are compared are measured growth
South Wales, over 30 years. The simulated values were calculated rates in plantations at the two sites, provided by
with 3-PG, using appropriate weather data and values for available
water in the root zone and fertility modifiers in Eq. Ž15.. Initial
forestry officers there.
stem populations were taken as 1500 hay1 . No fitting procedures The correspondence between measured and simu-
or adjustments were used. lated stem growth is clearly excellent, for both sites.
J.J. Landsberg, R.H. Waringr Forest Ecology and Management 95 (1997) 209–228 223

4. Discussion ratio is low, so that relatively small amounts of


precipitation improve the water balance—provides
4.1. Utilisable radiation feedback that ensures that the system generally be-
haves realistically.
One of the major simplifications in 3-PG, which
make it possible to keep the model relatively simple, 4.3. Constant PN r PG
is the use of modifiers to calculate utilisable radia-
tion, applied to monthly averages. There may be The use of a constant ratio Ž c pp . between P N and
some semantic objections to the term ‘utilisable’ PG eliminates the need for detailed calculations of
radiation, applied to the fraction of fp.a. that is respiration for the various tissues and components of
finally converted into dry mass by the canopy, on the trees, a procedure which, despite recent progress,
grounds that there are constraints other than those we would lead to large errors. The c pp values available
have allowed for—such as high light intensities in to us range from 0.37 to 0.5, but cluster round 0.45,
the top layers of canopies on bright days. However, suggesting that we can be relatively confident that
we regard the term as a useful and legitimate de- the value 0.45 " 0.05 that we have accepted can be
scription at the level of complexity with which we used with confidence. The errors in doing so are
are concerned. The procedure is well founded in almost certainly smaller than those that would result
physiological research and modelling using time steps from calculating respiration. Much more work is
of days Žsee Runyon et al., 1994; Waring et al., needed in this area, but we consider that this simpli-
1995., but will need to be tested against long-term, fication, which allows a major step forward in forest
detailed flux measurements, such as those now being modelling, is justifiable.
gathered at a number of sites ŽWofsy et al., 1993;
Goulden et al., 1996; Grace et al., 1996.. It can also 4.4. Carbon allocation
be evaluated in relation to models that use more
detailed light interception routines and physiological In later and more complex versions of this model
response algorithms. Such experimental measure- it may prove worthwhile Žgiven a sound basis for
ments and models will provide data that allow the doing so. to introduce accelerated leaf fall as a
range of variation in a c to be examined. consequence of long periods of drought, but we have
not done this.
4.2. Soil water modifier Given that we can calculate realistic values of
P N , the carbon allocation problem is one of the
The hydrology in this version of 3-PG is ex- major obstacles to the simulation of tree growth
tremely simple and may need to be reconsidered. patterns. We have solved it by assuming that the
The problem lies in the use of monthly water bal- allometric ratios obtained from measurements or de-
ances, which are inherently unsatisfactory. They as- structive harvesting of trees are genetically deter-
sume Žimplicitly. that rainfall is evenly distributed mined and that the trees tend to conserve these ratios
over the month, which is obviously not the case, so by conserving the ratios of the rates of change of
that soil recharge and depletion patterns are dis- foliage and stem mass ŽEq. Ž14..; i.e. allometric
torted. The problem could be solved by using daily ratios are the end result of carbon allocation up to
precipitation data and water balances, but this would the time they are measured. This is a robust assump-
be a move back to high data demands and increased tion; examination of the behaviour of Eq. Ž14. indi-
complexity—including the need to introduce inter- cates that it leads to variation with B of the type that
ception as a function of foliage mass, evaporation would be expected as trees age and stem diameters
from the canopy, etc. We do not currently consider increase. For any set of allometric equations where
this justified, and simply accept that the soil water n f - n s ŽEq. Ž9.. —which is the usual situation Žsee
modifier is rather crude. The fact that the soil water Table 2. —pf.s declines with B, so we do not end up
modifier affects the transpiration rates—reducing with enormous foliage mass, as would be the case if
them, for example, in months when the moisture partitioning was determined by the allometric equa-
224 J.J. Landsberg, R.H. Waringr Forest Ecology and Management 95 (1997) 209–228

tions used independently in their integral form. We maximum age of forest stands. Eq. Ž3., which serves
note that if this purpose in 3-PG, is entirely empirical and does
wf s a f W n f not invoke any particular mechanism, but it would be
an advantage, in relation to testing the equation, if
then the underlying mechanismŽs. could be unambigu-
dwf wf dW ously identified. We have indicated that, in our view,
s hf one of the most important of the various factors that
dt W dt
may be involved in the reduction in P N as stands
Similarly age is hydraulic conductivity, which feeds back to
dws ws dW stomata and hence photosynthesis Žsee Yoder et al.,
s hs 1994.. The results of Mencuccini and Grace Ž1996.
dt W dt
are consistent with this idea Žsee Fig. 3.. The matter
hence is currently receiving considerable attention and clar-
hf w f ification will, no doubt, emerge in due course.
pf .s s
hs ws
4.6. 3-PG as a practical forestry tool
showing explicitly that pf.s depends on hfrhs and
declines as ws increases. The excellent correspondence between simulated
Although we assume that allometric ratios are and measured cumulative stem biomass at Tumut
genetically determined and therefore reflect species and Haupapa ŽFig. 11., achieved simply by using the
differences Žsee Table 2 and Fig. 4., they are un- appropriate weather, soil moisture and fertility values
doubtedly modified by environmental conditions. We in 3-PG, gives confidence that the model is ready to
know that the rate of foliage production—and hence be used as a practical tool. We tested it against P.
n f in Eq. Ž9. —is lower for a given species in a radiata, which is, we acknowledge, a simple system,
harsh, as opposed to a mild environment. Similarly, because good growth measurements were available
stem biomass increases less with diameter in harsher from these plantations, which is seldom the case with
environments as a result of less height growth and native forests. However, we can, equally success-
more taper in stems ŽWaring, 1983; Waring and fully, simulate the biomass accumulation curves that
Schlesinger, 1985.. There is also evidence in both give the time course of current annual increment
angiosperms and gymnosperms that increases in arid- ŽCAI., developed by West and Mattay Ž1993.. In the
ity are associated with a reduction in specific leaf case of eucalypts the decrease in P N is much more
area ŽSpecht and Specht, 1989; Pierce et al., 1994.. rapid than in pines, and there has to be some element
As the environment becomes more arid, the leaf area of fitting by adjusting the constant in Eq. Ž3. that
is reduced but the carbon required to produce a unit controls the point at which P N begins to fall. It will
of leaf area more than doubles, affecting the growth be important, if 3-PG is to be established as a
rate and final tree size in predictable ways. practical tool, that many tests against measured stem
The procedure we have presented here is sensitive growth rates be carried out: the problem, in most
to differences in the parameters of Eq. Ž9., particu- cases, will be obtaining growth measurements good
larly because the key equation ŽEq. Ž14.. is a ratio of enough to do this. Observations of L) would also be
derivatives. This may impose stringent requirements of great value. Model outputs can be provided in
on experimental technique and field measurements, terms of changes in average B and it will be possible
but it does result in clear and testable hypotheses. to apply to these data standard distributions, such as
the Weibull distribution, to indicate the stem size
4.5. The age effect distribution that can be expected. Thinning proce-
dures have been written for the model, but these will
Allowing for the reduction in PN with stand age be described elsewhere.
is essential for any carbon balance model that is A matter of great interest, from the practical
intended to be run over periods approaching the forestry point of view, is the variation in wood
J.J. Landsberg, R.H. Waringr Forest Ecology and Management 95 (1997) 209–228 225

density. Physiological models of necessity work in us to develop a model which should be of value as a
terms of mass of carbon, and a value for wood research tool and has considerable potential value as
density of 500 kg my3 is generally accepted as a practical means of analysing forest growth and the
‘reasonable’. However, wood density may in fact reasons for its variations from place to place, and in
vary by a factor of more than two between species: time. The model also has potential as a tool for
Kingston and Risdon Ž1961. provide comprehensive ecological analysis. The carbon partitioning proce-
data for Australian trees, which show the density of dure, based on allometric ratios, provides interesting
P. radiata to be about 400 kg my3 and that of some insights into tree growth patterns and may be a
of the ‘Ironbark’ eucalyptus species as more than useful tool in genetic studies. The stem population
900 kg my3 . Such differences must reflect directly sub-model, based on the y3r2 power law, may also
the volume growth that can be achieved by different warrant further study in its own right.
species.

4.7. The use of remote sensing to parameterise the Acknowledgements


model
RHW acknowledges with thanks support from a
One of the primary aims in developing the 3-PG CSIRO McMaster Fellowship, support on sabbatical
model was to scale up predictions of stand growth to from his own department and a grant from the
region-wide estimates of PN so they could be com- Forestry and Forest Products Research and Develop-
pared with independent estimates derived from satel- ment Corporation, Australia. We thank Dr Heather
lites, which monitor seasonal and yearly variation in Keith for allometric data describing mature eucalypts
L) based on the ratio of the reflectances in the and for the information that carbon budgets of E.
infra-red and near infra-red wavebands. One of the pauciflora show the P N rPG ratio in those eucalypts
most commonly used indices, in this respect, is the to be about 0.4. We thank Dr Mike Ryan for unpub-
normalised difference vegetation index ŽNDVI; Run- lished information supporting the same P N rPG value
ning et al., 1989.. It is also possible to estimate fp.a. for a number of forest species and Robin Cromer for
directly from satellites and to derive estimates of the biomass data for young E. globulus plantations from
environmental constraints ŽGoward et al., 1994; which some of the allometric ratios in Table 2 were
Prince and Goward, 1995.. Remote sensing offers calculated, and we thank Dr. Mathew Williams for a
the only opportunity to observe changes in phenol- pre-publication copy of the paper ŽWilliams et al.,
ogy over large areas, and of indirectly obtaining an 1997. submitted to Ecological Applications. Prof.
estimate of soil water storage capacity in those areas Ross McMurtrie provided us with valuable com-
where water balance calculations have been made ments at the draft stage of this paper. Growth data
and correlated with seasonal reductions in L) . Coops for Tumut and Haupapa were provided by Mike
et al. Ž1997. have used a modified version of 3-PG Welch ŽState Forests of New South Wales. and Dr.
with remotely sensed measurements and have ob- David Whitehead ŽManaaki Whenua Landcare Re-
tained good correspondence between growth rates search, Lincoln, New Zealand..
derived from the average of forest plot measure-
ments over large areas Ž8 km = 8 km. and estimates
of mean annual growth made with this model. References

Aber, J.D., Federer, C.A., 1992. A generalized, lumped-parameter


model of photosynthesis, evapotranspiration and net primary
5. Conclusions production in temperate and boreal forest ecosystems. Oecolo-
gia 92, 463–474.
To conclude, we consider that our knowledge of Ågren, G.I., 1983. Nitrogen productivity of some conifers. Can. J.
For. Res. 13, 494–500.
the physiological processes underlying forest growth Ågren, G.I., Ingestad, T., 1987. Rootrshoot activity ratios as a
has reached the point where we can, with reasonable balance between nitrogen productivity and photosynthesis.
confidence, make simplifications that have allowed Plant, Cell Environ. 10, 579–586.
226 J.J. Landsberg, R.H. Waringr Forest Ecology and Management 95 (1997) 209–228

Beets, P.N., Madgwick, H.A.I., 1988. Above-ground dry matter turbed tropical forest in south-west Amazonia. Global Change
and nutrient content of Pinus radiata as affected by lupins, Biol. 1, 1–12.
fertiliser, thinning and stand age. N.Z. J. For. 18, 43–64. Granier, A., Biron, P., Brea, ´ N., Pontailler, J.-Y., Saugier, B.,
Beets, P.N., Whitehead, D., 1996. Carbon partitioning in Pinus 1996. Transpiration of trees and forest stands: short and
radiata in relation to foliage nitrogen status. Tree Physiol. 16, long-term monitoring using sapflow methods. Global Change
131–138. Biol. 2, 265–274.
Benecke, U., Evans, G., 1987. Growth and water use in Nothofa- Johnson, D.W., Linberg, S.E. Žeds.., 1992. Atmospheric Deposi-
gus truncata ŽHead beech. in temperate hill country, Nelson, tion and Forest Nutrient Cycling. Springer, New York, 707 pp.
New Zealand. In: Hanxi, Y., Jeffers, J.N.R., Ward, P.A. Žeds.., Kelliher, F.M., Leuning, R., Schulze, E.-D., 1993. Evaporation
The Temperate Forest Ecosystem. Symp. 20. Institute of Ter- and canopy characteristics of coniferous forest and grassland.
restrial Ecology, Grange-over-Sands, UK, pp. 131–140. Oecologia 95, 153–163.
Cannell, M.G.R., Dewar, R.C., 1994. Carbon allocation in trees: a Kelliher, F.M., Leuning, R., Raupach, M.R., Schulze, E.-D., 1995.
review of concepts for modelling. Adv. Ecol. Res. 25, 60–140. Maximum conductances for evaporation from global vegeta-
Comins, H.N., McMurtrie, R.E., 1993. Long-term response of tion types. Agric. For. Meteorol. 73, 1–16.
nutrient-limited forests to CO 2 enrichment: equilibrium behav- Kingston, R.S.T., Risdon, C.J.E., 1961. Shrinkage and density of
ior of plant–soil models. Ecol. Appl. 3, 666–681. Australian and other south-west Pacific woods. Div. For. Prod.
Coops, N.C., Waring, R.H., Landsberg, J.J., 1997. Assessing Tech. Pap. 13, CSIRO, Melbourne, Australia, 65 pp.
forest productivity in Australia and New Zealand using a ¨
Korner, C., 1993. Leaf diffusive conductances in the major vege-
physiologically-based model driven with averaged monthly tation types of the globe. In: Schulze, E.-D., Caldwell, M.M.
weather data and satellite-derived estimates of canopy photo- Žeds.., Ecophysiology of Photosynthesis. Ecological Studies,
synthetic capacity. For. Ecol. Manage. Žsubmitted.. Vol. 100. Springer, Berlin, pp. 463–490.
Denmead, O.T., Shaw, R.H., 1961. Availability of soil water to Kuuluvaineen, T., 1991. Long-term development of needle mass,
plants as affected by soil moisture content and meteorological radiation interception and stemwood production in naturally
conditions. Agron. J. 54, 385–390. regenerated Pinus sylÕestris stands on Empetrum – Vaccinium
Drew, T.J., Flewelling, J.W., 1977. Some recent Japanese theories site type in the northern boreal zone in Finland: an analysis
of yield–density relationships and their applications to Mon- based on an empirical study and simulation. For. Ecol. Man-
terey pine plantations. For. Sci. 23, 517–534. age. 46, 103–122.
Dunin, F.X., McIlroy, I.C., O’Loughlin, E.M., 1985. A lysimeter Landsberg, J.J., 1986. Physiological Ecology of Forest Production.
characterization of evaporation by Eucalypt and its representa- Academic Press, Sydney, 198 pp.
tiveness for the local environment. In: Hutchinson, B.A., Landsberg, J.J., Gower, S.T., 1997. Applications of Physiological
Hicks, B.B. Žeds.., The Forest–Atmosphere Interaction. Rei- Ecology to Forest Management. Academic Press, San Diego,
del, Dordrecht, pp. 271–291. CA, 354 pp.
Dye, P.J., Olbrich, B.W., 1993. Estimating transpiration from Landsberg, J.J., Hingston, F.J., 1996. Evaluating a simple radia-
6-year-old Eucalyptus grandis trees: development of a canopy tionrdry matter conversion model using data from Eucalyptus
conductance model and comparison with independent sap flux globulus plantations in Western Australia. Tree Physiol. 16,
measurements. Plant, Cell Environ. 16, 45–53. 801–808.
Gholz, H.L., Grier, C.C., Campbell, A.G., Brown, A.T., 1979. Landsberg, J.J., Prince, S.D., Jarvis, P.G., McMurtrie, R.E., Lux-
Equations for estimating biomass and leaf area of plants in the moore, R., Medlyn, B.E., 1996. Energy conversion and use in
Pacific Northwest. Res. Pap. 41, Forestry Research Labora- forests: an analysis of forest production in terms of radiation
tory, Oregon State University, School of Forestry, Corvallis, utilisation efficiency. In: Gholz, H.L., Nakane, K., Shimoda,
38 pp. H. Žeds.., The Use of Remote Sensing in the Modeling of
Goulden, M.L., Munger, J.W., Fan, S.-M., Daube, B.C., Wofsy, Forest Productivity. Kluwer Academic, Dordrecht, Nether-
S.C., 1996. Measurements of carbon sequestration by long-term lands, pp. 273–298.
eddy covariance: methods and a critical evaluation of accu- Leuning, R., 1995. A critical appraisal of a combined stomatal–
racy. Global Change Biol. 2, 169–182. photosynthesis model for C 3 plants. Plant, Cell Environ. 18,
Goward, S.N., Waring, R.H., Dye, D.G., Yang, J., 1994. Ecologi- 339–355.
cal remote sensing at OTTER: satellite macroscale observa- Linder, S., Benson, M.L., Myers, B.J., Raison, R.J., 1987. Canopy
tions. Ecol. Appl. 4, 322–343. dynamics and growth of Pinus radiata. I. Effects of irrigation
Gower, S.T., Gholz, H.L., Nakane, K., Baldwin, V.C., 1994. and fertilization during a drought. Can. J. For. Res. 17,
Production and carbon allocation patterns of pine forests. Ecol. 1157–1165.
Bull. 43, 115–135. McMurtrie, R.E., 1985. Forest productivity in relation to carbon
Gower, S.T., McMurtrie, R.E., Murty, D., 1996. Aboveground net partitioning and nutrient cycling: a mathematical model. In:
primary production decline with stand age: potential causes. Cannell, M.G.R., Jackson, J.E. Žeds.., Trees as Crop Plants.
Trends Ecol. Evol., 11, 378–382. Institute of Terrestrial Ecology, Natural Environment Research
Grace, J., Lloyd, J., McIntyre, J., Miranda, A., Meir, P., Miranda, Council, UK, pp. 194–207.
H., Moncrieff, J., Massheder, J., Wright, I., Gash, J., 1996. McMurtrie, R.E., Landsberg, J.J., 1992. Using a simulation model
Fluxes of carbon dioxide and water vapour over an undis- to evaluate the effects of water and nutrients on the growth
J.J. Landsberg, R.H. Waringr Forest Ecology and Management 95 (1997) 209–228 227

and carbon partitioning of Pinus radiata. For. Ecol. Manage. Ryan, M.G., 1991. A simple method for estimating gross carbon
52, 243–260. budgets for vegetation in forest ecosystems. Tree Physiol. 9,
McMurtrie, R.E., Wolf, L., 1983. Above- and below-ground 255–266.
growth of forest stands: a carbon budget model. Ann. Bot. 52, Ryan, M.G., Gower, S.T., Hubbard, R.M., Waring, R.H., Gholz,
437–448. H.L., Cropper Jr., W.P., Running, S.W., 1995. Woody tissue
McMurtrie, R.E., Comins, H.N., Kirschbaum, M.U.F., Wang, maintenance respiration of four conifers in contrasting cli-
Y.P., 1992. Modifying existing forest growth models to take mates. Oecologia 101, 133–140.
account of effects of elevated CO 2 . Aust. J. Bot. 40, 657–77. Ryan, M.G., Yoder, B.J., 1997. Hydraulic limits to tree height and
McMurtrie, R.E., Gholz, H.L., Linder, S., Gower, S.T., 1994. tree growth. BioScience, 47, 235–242.
Climatic factors controlling the productivity of pine stands: a Ryan, M.G., Hubbard, R.M., Pongracic, S., Raison, R.J., McMur-
model-based analysis. Ecol. Bull. 43, 173–188. trie, R.E., 1996a. Foliage, fine-root, woody-tissue and stand
Mencuccini, M., Grace, J., 1996. Hydraulic conductance, light respiration in Pinus radiata in relation to nitrogen status. Tree
interception and needle nutrient concentration in Scots pine Physiol. 16, 333–343.
stands and their relation with net primary productivity. Tree Ryan, M.G., Binkley, D., Fownes, J.H., 1996b. Age-related de-
Physiol. 16, 459–468. cline in forest productivity: pattern and process. Adv. Ecol.
Monteith, J.L., 1977. Climate and the efficiency of crop produc- Res., 27, 213–262.
tion in Britain. Phil. Trans. R. Soc., Ser. B 281, 277–294. Santantonio, D., 1989. Dry-matter partitioning and fine-root pro-
Murty, D., McMurtrie, R.E., Ryan, M.G., 1996. Declining forest duction in forests—new approaches to a difficult problem. In:
productivity in ageing forest stands: a modelling analysis of Pereira, J.S., Landsberg, J.J. Žeds.., Biomass Production by
alternative hypotheses. Tree Physiol. 16, 187–200. Fast-Growing Trees. Kluwer Academic, Dordrecht, Nether-
Pastor, J., Bockheim, J.G., 1981. Biomass and production of an lands, pp. 57–72.
aspen–mixed hardwood–spodosol ecosystem in northern Wis- Schulze, E.-D., Fuchs, M.I., Fuchs, M., 1977. Spatial distribution
consin. Can. J. For. Res. 11, 132–138. of photosynthetic capacity and performance in a mountain
Pastor, J., Aber, J.D., Melillo, J.M., 1984. Biomass prediction spruce forest of northern Germany. Oecologia 29, 4361.
using generalized allometric regressions for some northeast Snowdon, P., Benson, M.L., 1992. Effects of combinations of
tree species. For. Ecol. Manage. 7, 265–274. irrigation and fertilisation on the growth and above-ground
Pearson, J.A., Fahey, T.J., Knight, D.H., 1984. Biomass and leaf biomass production of Pinus radiata. For. Ecol. Manage. 52,
area in contrasting lodgepole pine forests. Can. J. For. Res. 14, 87–116.
259–265. Specht, R.L., Specht, A., 1989. Canopy structure in Eucalyptus-
Pierce, L.L., Running, S.W., Walker, J., 1994. Regional scale dominated communities in Australia along climatic gradients.
relationships of leaf area index to specific leaf area and leaf Acta Oecol. 10, 191–213.
nitrogen content. Ecol. Appl. 4, 313–321. Thornley, J.H.M., 1972a. A model to describe the partitioning of
Pook, E.W., 1986. Canopy dynamics of Eucalyptus maculata photosynthate during vegetative plant growth. Ann. Bot. 36,
Hook. IV. Contrasting responses to two severe droughts. Aust. 419–430.
J. Bot. 34, 1–14. Thornley, J.H.M., 1972b. A balanced quantitative model for
Prince, S.D., Goward, S.N., 1995. Global primary production: a root:shoot ratios in vegetative plants. Ann. Bot. 36, 431–441.
remote sensing approach. J. Biogeogr. 22, 815–835. Tyree, M.T., Sperry, J.S., 1988. Do woody plants operate near the
Running, S.W., Coughlan, J.C., 1988. A general model of forest point of catastrophic xylem disfunction caused by dynamic
ecosystem processes for regional applications. I. Hydrologic water stress? Plant Physiol. 88, 574–580.
balance, canopy gas exchange and primary production pro- Wang, Y.-P., McMurtrie, R.E., Landsberg, J.J., 1992. Modelling
cesses. Ecol. Modelling 42, 125–154. canopy photosynthetic productivity. In: Baker, N.R., Thomas,
Running, S.W., Gower, S.T., 1991. FOREST-BGC, a general H. Žeds.., Crop Photosynthesis: Spatial and Temporal Determi-
model of forest ecosystem processes for regional applications. nants. Elsevier, New York, pp. 43–67.
II. Dynamic carbon allocation and nitrogen budgets. Tree Waring, R.H., 1983. Estimating forest growth and efficiency in
Physiol. 9, 147–160. relation to canopy leaf area. Adv. Ecol. Res. 13, 327–354.
Running, S.W., Hunt, E.R., 1993. Generalization of a forest Waring, R.H., Major, J., 1964. Some vegetation of the California
ecosystem process model for other biomes, BIOME-BGC, and coastal redwood region in relation to gradients of moisture,
an application for global scale models. In: Ehleringer, J.R., nutrients, light, and temperature. Ecol. Monogr. 34, 167–215.
Field, C.B. Žeds.., Scaling Physiological Processes: Leaf to Waring, R.H., Schlesinger, W.H., 1985. Forest Ecosystems: Con-
Globe. Academic Press, San Diego, pp. 141–158. cepts and Management. Academic Press, Orlando, FL, 340 pp.
Running, S.W., Nemani, R.R., Peterson, D.L., Band, L.E., Potts, Waring, R.H., Law, B.E., Goulden, M.L., Bassow, S.L., Mc-
D.F., Pierce, L.L., Spanner, M.A., 1989. Mapping regional Creight, R.W., Wofsy, S.C., Bazzaz, F.A., 1995. Scaling gross
forest evapotranspiration and photosynthesis by coupling satel- ecosystem production at Harvard Forest with remote sensing:
lite data with ecosystem simulation. Ecology 70, 1090–1101. a comparison of estimates from a constrained quantum-use
Runyon, J., Waring, R.H., Goward, S.N., Welles, J.M., 1994. efficiency model and eddy correlation. Plant, Cell Environ. 18,
Environmental limits on net primary production and light-use 1201–1213.
efficiency across the Oregon transect. Ecol. Appl. 4, 226–237. Weinstein, D.A., Beloin, R.M., Yanai, R.D., 1991. Modeling
228 J.J. Landsberg, R.H. Waringr Forest Ecology and Management 95 (1997) 209–228

changes in red spruce carbon balance and allocation in re- Williams, M., Rastetter, E.B., Fernandes, D.N., Goulden, M.L.,
sponse to interacting ozone and nutrient stresses. Tree Physiol. Shaver, G.R., 1997. Predicting gross primary productivity in
9, 127–146. terrestrial ecosystems. Ecol. Appl., in press.
West, P.W., Mattay, J.P., 1993. Yield prediction models and Wofsy, S.C., Goulden, M.L., Munger, J.W., Fan, S.-M., Bakwin,
comparative growth rates for six eucalypt species. Aust. For. P.S., Daube, B.C., Bassow, S.L., Bazzaz, F.A., 1993. Net
56, 211–225. exchange of CO 2 in a mid-latitude forest. Science 260, 1314–
White, J., 1981. The allometric interpretation of the self-thinning 1317.
rule. J. Theor. Biol. 89, 475–500. Yoder, B.J., Waring, R.H., 1994. The normalized difference vege-
White, J.D., Running, S.W., 1994. Testing scale dependent as- tation index of small Douglas-fir canopies with varying
sumptions in regional ecosystem simulations. J. Veg. Sci. 5, chlorophyll concentrations. Remote Sens. Environ. 49, 81–91.
687–702. Yoder, B.J., Ryan, M.G., Waring, R.H., Schoettle, A.W., Kauf-
Williams, J., Prebble, R.E., Williams, W.T., Hignett, C.T., 1983. mann, M.R., 1994. Evidence of reduced photosynthetic rates
The influence of texture, structure and clay mineralogy on the in old trees. For. Sci. 40, 513–427.
soil moisture characteristic. Aust. J. Soil 21, 15–32.

Вам также может понравиться