Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

ECOHYDROLOGY

Ecohydrol. 9, 610–630 (2016)


Published online 7 September 2015 in Wiley Online Library
(wileyonlinelibrary.com) DOI: 10.1002/eco.1661

Improving the ability of 3-PG to model the water balance of


forest plantations in contrasting environments
Auro C. Almeida1* and Peter J. Sands2
1
CSIRO, Land and Water, Private Box 12, Hobart, Tasmania, Australia, 7001
2
39 Oakleigh Av, Taroona, Tasmania, Australia, 7053

ABSTRACT
Because water availability is a primary factor that influences growth of forest plantations, quantifying plantation water use is an
important part of the decision-making process when selecting new plantation areas. The 3-PG model (Landsberg and Waring,
Forest Ecology and Management 95: 209–228, 1997) has been used successfully to model forest growth for several species.
However, some studies have demonstrated limitations in 3-PG’s ability to accurately model the soil and plant water balance. This
paper addresses these limitations through the development and validation of a modified water balance submodel that models
canopy interception, soil evaporation, tree transpiration, drainage, run-off, soil water content and available soil water in the root
and non-root zone. The new submodel was tested and validated using detailed, long-term observed growth and water balance
data from contrasting environments in Australia and Brazil and showed improved estimation of available soil water, even when
only monthly weather data were available. However, the new submodel requires more soil texture-dependent parameters than the
original submodel. Sensitivity analysis of these new parameters shows that values of soil field capacity, wilting point and soil
depth may have strong impacts on the prediction of run-off, soil water content and available soil water. Copyright © 2015 John
Wiley & Sons, Ltd.

KEY WORDS 3-PG model; water balance; plantation; forest growth; water use; available soil water

Received 24 May 2015; Revised 19 June 2015; Accepted 21 June 2015

INTRODUCTION 3-PG includes a simple soil water balance submodel


comprising a single soil layer, with evapotranspiration (ET)
Globally, the area of forest plantations is increasing, mainly
driven by stand-level canopy leaf area index (LAI) and
in drier regions where there is competition for water
canopy conductance. This water balance submodel has
resources (Winjum and Schroeder, 1997; Almeida et al.,
been criticized on a number of counts. Firstly, water stress
2010). In order to optimally manage these plantations,
is based on the ASW of the entire soil profile, whereas it is
models that accurately predict their growth and water use
only the water content of the root zone that directly affects
by coupling stand growth and forest soil water balance are
plant water status (Whitehead and Beadle, 2004). Second-
required. Examples are Forest BGC (Running and
ly, tree transpiration and soil evaporation are combined as a
Coughlan, 1988), CABALA (Battaglia et al., 2004) and
single flux of water, whereas in reality, they are separate
3-PG (Landsberg and Waring, 1997; Sands and Landsberg,
fluxes that depend on environmental and stand factors in
2002; Landsberg and Sands, 2010).
different ways. Thirdly, the dynamics of soil water balance
3-PG is a process-based forest growth model widely used
can occur on a much shorter timescale than a month, so the
as a research and management tool (Almeida et al., 2004a;
monthly time step fails to account for these changes (e.g.
Landsberg and Sands, 2010, Section 9.4) and has been
Paul et al., 2006; Almeida et al., 2007).
extensively applied to plantations or even-aged, relatively
In this paper, we present a more detailed water balance
homogeneous forests. It operates on a monthly time step and
submodel that improves the ability of 3-PG to predict the
predicts stem, foliage and root biomass, number of trees per
stand-level forest soil water balance, while maintaining the
hectare, and available soil water (ASW) in response to
simplicity of 3-PG and its ability to accurately predict forest
changing weather conditions and management interventions
growth. To address the aforementioned criticisms and
such as irrigation, thinning and fertilization.
deficiencies, the new submodel uses water availability in
the root zone to determine the effects of soil water stress,
*Correspondence to: Auro C. Almeida, CSIRO, Land and Water, Private
Box 12, Hobart, Tasmania, Australia, 7001. considers transpiration and soil water evaporation as distinct
E-mail: Auro.Almeida@csiro.au processes that depend on stand characteristics and environ-

Copyright © 2015 John Wiley & Sons, Ltd.


MODEL WATER BALANCE OF FOREST PLANTATIONS IN CONTRASTING ENVIRONMENTS 611

mental factors in different ways and allows the use of shorter management decision-making (Landsberg, 2003; Almeida
time steps than a month, including a daily time step. et al., 2004a).
However, this approach requires more soil texture- A full description of 3-PG is given by Landsberg and
dependent parameters than does the original submodel. Sands (2010), which we abbreviate in this paper to L&S.
Also, rigorous validation of the model requires long-term Here, we focus on the water balance submodel. The
observed data on the separate fluxes in the water balance, original water balance submodel in 3-PG, called WB1 in
which may not always be available. We used data from this paper, assumes the soil profile is homogeneous and of
plantations grown in contrasting environments that together a fixed depth and that roots always have access to all water
provide observed data for all the fluxes in the water stored in the profile. The plant ASW in the full profile is the
balance. To test the model for different species and focus of the model and mediates the effects of water stress
contrasting soils, climates and water availability, we used on growth. Water availability in the profile is specified by
detailed forest growth and water balance measurements the maximum ASW capacity and two parameters cθ and nθ
from Eucalyptus grandis hybrids grown on the east coast of that characterize the soil water-dependent growth modifier
Brazil (Soares and Almeida, 2001; Almeida et al., 2004b, fθ. WB1 does not predict actual soil water content (SWC).
2007) and from Eucalyptus globulus plantations in Victoria The monthly change in ASW is the monthly rainfall less the
and South Australia (Benyon et al., 2006) and in Western sum of a rainfall interception loss, surface run-off and stand
Australia (WA) (Mendham et al., 2011). transpiration ET (which combines canopy transpiration and
We test and compare the performance of the original soil evaporation). A fraction of rainfall is intercepted by the
and new water balance submodels using three different foliage and subsequently returned to the atmosphere (through
time steps: (i) monthly; (ii) ‘pseudo-daily’, where the time evaporation). The remaining rainfall is added to the ASW
step is the average time between precipitation events; and pool, and any excess over the maximum ASW capacity of the
(iii) daily. Sensitivity analyses were conducted to quantify profile is deemed lost as surface run-off and deep drainage.
the sensitivity of water balance and predicted tree growth Stand transpiration is determined using an expression based on
to the new soil-specific parameters. the Penman–Monteith equation. This requires a stand-level
canopy conductance gC, which increases linearly with
increasing canopy LAI up to some maximum value. When
MODEL DESCRIPTION LAI is zero or small, gC is non-zero because ET includes soil
evaporation. Conductance is also modified by functions of
Overview of 3-PG temperature, vapour pressure deficit (VPD), ASW and
3-PG is driven by monthly climatic data and requires atmospheric CO2 (L&S, Section 9.2.7; Almeida et al., 2009).
knowledge of the local soil texture, soil water holding
capacity and soil fertility. Simple relationships describe the
processes underpinning tree growth, mortality and water use The new water balance submodel
on a monthly timescale. Many of the parameters in these The three deficiencies of WB1 we address here are as follows:
relationships are species specific. Values are also required at (i) not all the water in the soil profile is available for
some initial age for stand stocking, the stem (including transpiration until roots have fully occupied the profile; (ii)
bark and branches), foliage and root biomass per hectare, soil evaporation and canopy transpiration depend on stand
and the ASW. characteristics and environmental factors in different ways
The parameterization of 3-PG has been described and and should not be aggregated; and (iii) the monthly time step
illustrated for a range of individual species or genotypes within ignores the fact that rainfall occurs in discrete events, whose
a species. 3-PG is used worldwide for different tree species temporal distribution affects the soil water balance. The new
such as E. globulus (Sands and Landsberg, 2002; Fontes et al., water balance model, called WB2 in this paper, addresses
2006; Miehle et al., 2009; Smettem et al., 2013), Eucalyptus these issues by (i) subdividing the profile into an upper
nitens (Rodriguez et al., 2009; Pérez-Cruzado et al., 2011; horizontal zone containing roots and a lower root-free zone,
Gonzalez-Garcia et al., 2015), E. grandis (Esprey et al., 2004; with the effects of water stress based only on the water content
Stape et al., 2004; Almeida et al., 2004b, 2007), Eucalyptus of the root zone, (ii) modelling canopy transpiration and soil
cladocalyx (Paul et al., 2007), Pinus patula (Dye, 2001), Pinus evaporation as distinct processes and (iii) explicitly consid-
radiata (Coops, 1999; Rodríguez et al., 2002; Feikema et al., ering time steps that are less than a month. WB2 predicts the
2010) and more recently to mixed species (Forrester and Tang, SWC of the profile, and ASW is determined from this, the
2015). The model has been applied at stand and spatial scales current root depth and texture-dependent soil characteristics.
(Coops and Waring, 2001; Tickle et al., 2001; Sands, 2004; The principles of WB2 are described here; mathematical
Almeida et al., 2010) for different purposes, e.g. to estimate details are in Appendix A.
potential productivity or the impact of climate and silvicultural The hydrological balance in WB2 is as described in L&S
managements on forest yield and as a research tool for forest (Section 7.1), except that recharge from a water table and

Copyright © 2015 John Wiley & Sons, Ltd. Ecohydrol. 9, 610–630 (2016)
612 A. C. ALMEIDA AND P. J. SANDS

lateral flow are not considered [Equation (A2)]. Input of water SWCs. The recharge flux is assumed to be proportional to
is assumed to be from rainfall, and possibly irrigation. the difference between the volumetric SWCs (θ) of the two
Contributions from snow are not explicitly considered. zones (Appendix A.7), and water flows from the region
However, such contributions could be included by using a with the higher θ to that with the lower θ.
simple model of snow melt to allocate additional daily WB2 requires several texture-dependent soil character-
amounts of equivalent rain. istics. These include the volumetric SWCs of the soil at
Rainfall is partially lost through canopy interception and wilting point, field capacity (above which free drainage
subsequent evaporation, and the remaining throughfall is occurs) and saturation (θwp, θfc and θsat; m3 m3).
added to the root zone. It is assumed that all input water comes Additional texture-dependent characteristics are the param-
from rainfall, possibly including irrigation. Snow contribu- eters cθ and nθ that characterize fθ, the rate constant (kDrain,
tions have not been modelled because our experimental sites day1) for drainage of water through the soil and a measure
are not subjected to snow. The fraction of rainfall lost by (kS, m day1) of soil hydraulic conductivity determining
interception is determined by canopy LAI up to a maximum the rate of root-zone recharge. These can be assigned by
fraction, which is a parameter of the species [Equation (A5)]. direct measurement, or on the basis of published soils data
Once the profile is saturated, any further throughfall is lost as (Williams et al., 1983; Saxton and Rawls, 2006). The
run-off. Free drainage occurs from the root zone into the root- fraction pstones of soil volume occupied by stones is also
free zone, and from the root-free zone out of the bottom of the taken in account.
profile. Soil evaporation and transpiration remove water from
the root zone, and recharge of the root zone can occur from the Transpiration and soil evaporation. Canopy transpiration
root-free zone through hydraulic pressure. and soil evaporation are modelled separately in WB2
(Appendix A.4). Transpiration is modelled as in WB1,
Root and root-free zones. The soil profile in WB2 is still using the Penman–Monteith equation, but because it now
homogeneous but is divided into a root zone, from which excludes soil evaporation, gC is the conductance of the
water is lost by transpiration and soil evaporation, and a canopy alone and increases from zero as LAI increases
non-root zone from which water cannot be immediately from zero, and transpiration is driven by the radiation
accessed by trees. The soil water-dependent growth absorbed by the canopy, not incident upon it.
modifier (fθ) is defined as in WB1 but is now based on Soil evaporation is implicitly confined to a shallow
the water content of the root zone only. upper soil layer that acts as a barrier to further evaporation
The ASW, i.e. the amount of water above wilting point, is as the soil dries. Soil evaporation rate (eS) is high when the
calculated only for the root zone and is used to determine the soil surface is wet after rain and declines as it dries out.
effects of water stress on growth and canopy conductance. This is modelled (Appendix A.5) using the so-called two-
This is performed using a growth modifier fθ based on phase Ritchie (1972) model. In Phase 1 (after a wetting
relative ASW θr in the root zone [Equation (A4)]. This event), eS is the wet-surface rate and determined using a
modifier varies from zero when θr = 0 to 1 when θr = 1 and is Penman–Monteith equation. Phase 2 commences once
applied to both growth and canopy conductance. accumulated evaporation exceeds an amount ES1, and eS is
The depth zR (mm) of the root zone is assumed to be then assumed to decline hyperbolically with increasing
proportional to current root biomass, up to some site- evaporation accumulated above that in Phase 1. The rate of
specific maximum depth zRx (Appendix A.7). Below this decline is described by a parameter (ES2), which is the
limit, each kilogram of root is assumed to exploit a volume amount of evaporation in Phase 2 required to reduce eS by
σ zR (m3 kg1) of soil, where σ zR is called the specific root 50%. The equations are Equations (A9) and (A10). ES1 and
volume. As root biomass increases, the volume of the root ES2 are potentially soil texture dependent. WB2 takes into
zone increases at the expense of the root-free volume, and account the difference between evaporation in full sun in a
the water contents of these two volumes are adjusted to partially open canopy and in the shade under the canopy.
take this change into account. Provision is made for the
possibility that SWC is observed only down to some depth Pseudo-daily and daily time steps. Two approaches can be
zO, less than the profile depth zS, but the full profile is used to improve the time resolution of the water balance
always used when determining the water balance. submodel. The first is to use daily weather data and to run
Drainage from the root zone into the root-free zone and both the growth and water balance submodels of 3-PG with a
from the latter out of the bottom of the profile is daily time step. The second approach uses traditional
considered, and the rate of free drainage is proportional monthly 3-PG input data provided the average number nR
to the amount of water above field capacity (Appendix of days on which rain falls in each month is available. If R is
A.6). Recharge of the root zone is based on a simple the total rain for the month, the month is divided into nR
representation of the movement of water across a boundary periods of equal length, each with an amount R/nR of rain.
between two regions of soil with different volumetric This rain is assumed to fall on the first day of each period,

Copyright © 2015 John Wiley & Sons, Ltd. Ecohydrol. 9, 610–630 (2016)
MODEL WATER BALANCE OF FOREST PLANTATIONS IN CONTRASTING ENVIRONMENTS 613

with the remaining days dry. These nR periods are then used weather data of selected study sites in WA, Green Triangle and
as time steps with 3-PG applied to each period. We call this Brazil and clearly demonstrates the strong variability between
approach ‘pseudo-daily’, and it can be used equally with the sites mainly in terms of rainfall distribution and air temperature.
original or new soil water balance submodels. Simple Further details are given in the following.
expressions can be written for each water balance compo- Procedures for the calibration and validation of the
nent over each pseudo-daily time period (Appendix A). original 3-PG model for E. grandis and E. globulus have
been published elsewhere (Sands and Landsberg, 2002;
Esprey et al., 2004; Almeida et al., 2004b).
METHODS
Aracruz site
Sites and species The Aracruz site is in an experimental catchment that is well
Contrasting climatic conditions, soils and species are desirable documented and described (Soares and Almeida, 2001;
to test and validate the performance of the model. We used data Almeida et al., 2004b, 2007). Plantation growth and water
from plantations of E. grandis hybrids grown at Aracruz on the balance were intensively monitored for a period of more than
east coast of Brazil and from E. globulus grown at contrasting 7 years, and rainfall, soil moisture, groundwater level and run-
sites in Southern Australia. The E. globulus data come from off were monitored automatically and continuously. Tree
eight sites in the Green Triangle area of Victoria and South transpiration (Mielke et al., 1999), canopy rainfall intercep-
Australia and from two sites in WA. All sites had been tion and destructive biomass samples (stem, bark, branch,
extensively monitored for growth and water use over a long foliage and roots) were obtained from intensive seasonal or
period. Table I lists the experimental sites, their location and annual campaigns. Tree growth was measured monthly in
main characteristics. Figure 1 shows the average monthly permanent sample plots (Almeida et al., 2004a).

Table I. Details of experimental sites.

Mean annual Age of Planted Water


precipitation measurement Planted stocking table depth
Site name Species Lat. Long. (mm year1) (years) date (tree ha1) Soil texture (m) Sourcea

Aracruz E. grandis 19·861 40·211 1250 1–7 2/1997 1111 Clay loam 15.0 1, 2
hybrid
Bessie E. globulus 38·158 141·818 786 6–9 7/1998 1100 Loamy sand/clay loam 6·0 3
DPI E. globulus 37·847 142·076 679 8–11 7/1996 950 Clay loam/clay 10·0 3
Gummy E. globulus 37·737 140·782 710 4–9 7/1996 1125 Loamy sand/clay 4·3 3
Jack E. globulus 37·373 140·456 664 3–7 7/1998 1175 Loamy sand/clay 3·3 3
Jasper E. globulus 37·836 141·620 706 7–9 7/1998 750 Sandy loam/clay 20·0 3
Jill E. globulus 37·373 140·456 643 4–7 7/1998 1200 Loamy sand 10·5 3
Padthaway E. globulus 36·541 140·652 503 4–6 7/1995 1025 Loamy sand 15·9 3
Vikki E. globulus 38·104 142·146 736 4–9 8/1998 925 Clay loam/clay 8·3 3
Wellstead E. globulus 34·580 118·550 591 2–8 7/1996 1131 Sand n/a 4, 5
Scott River E. globulus 34·260 115·520 984 2–8 7/1996 1145 Sandy clay loam n/a 4, 5
a
Sources: 1, Almeida et al. (2004a); 2, Soares and Almeida (2001); 3, Benyon et al. (2006); 4, White et al. (2009); 5, Mendham et al. (2011).

Figure 1. A comparison of the climate of the three experimental sites: (a) Aracruz in Brazil, (b) DPI in the Green Triangle and (c) Scott River in Western
Australia (S. River). Data shown are monthly precipitation (bars, mm), monthly average maximum (■) and minimum (●) temperature (°C) and monthly
2 1
average global radiation (♦, MJ m day ).

Copyright © 2015 John Wiley & Sons, Ltd. Ecohydrol. 9, 610–630 (2016)
614 A. C. ALMEIDA AND P. J. SANDS

The catchment is located on the coast of Espírito Santo below the surface. Measurements of rainfall, canopy
state. It has an area of 286 ha with 190 ha of E. grandis interception, soil evaporation, tree transpiration and changes
hybrid plantations at spacing of 3 m × 3 m located in the flat in SWC (Benyon et al., 2008) were made approximately
part of the catchment favouring water infiltration rather than monthly. SWC was measured at 0·3-m intervals of depth
run-off and 86 ha of preserved Atlantic rainforest in the using a neutron probe in tubes to a maximum depth of
valley and riparian zone. Mean annual precipitation is 5·85 m, and the changes in soil moisture between measure-
1250 mm, average temperature is 23·5 °C and solar radiation ments were evaluated for a period of at least 2 years. Details
averages 17 MJ m2 day1. Meteorological data were col- of the measurements are given by Benyon et al. (2008).
lected in three automatic weather stations at 15-min, hourly The FR in the Green Triangle and WA sites was based
and daily intervals. Daylight VPD was calculated as in on a decision tree using soil type and previous land use
Almeida and Landsberg (2003). Daily climate data were (Paul et al., 2007). Site FR was therefore generally
used in the Aracruz site in order to compare the model assumed to be between 0·7 and 0·9, or between 0·4 and
performance using daily or monthly time steps. 0·6 if soils were particularly infertile or growth was
The soil at the Aracruz site is classified as Agrisol with a impeded by salinity, shallow soil depth or poor drainage.
predominant clay-loam texture; a detailed soil description
is in Almeida et al. (2004b). Soil moisture under the WA sites
plantation was measured at weekly or fortnightly intervals In this study, we selected two contrasting sites in WA from
from age 2 to 7·5 years using a neutron probe in access those described in detail by White et al. (2009) and
tubes at 0·2-m intervals of depth to a maximum of 2·6 m Mendham et al. (2011): Scott River and Wellstead. At
(Almeida et al., 2007; Smethurst et al., 2015). These data both, initial planting density was 1250 trees per hectare,
were calibrated against volumetric SWC of samples and growth, LAI and soil moisture had been measured.
collected in the dry and wet seasons at each site, and the Mean annual rainfall was 591 and 984 mm for these sites,
results presented as total SWC. ASW was calculated as the respectively. The soils in the Scott River and Wellstead
difference between measured SWC and the minimum SWC sites are deep (up to 8 and 18 m, respectively) with a sandy
over the measurement period. Fertility rating (FR) was A-horizon and show evidence of laterite in the top 2 m with
assigned as in Almeida et al. (2010), and the sites from clay sub-soils (White et al., 2009). Measurements in these
which data were collected were well managed (good weed plots were from September 1998 to September 2004.
control and fertilizer applied).

Green Triangle sites Model calibration and validation

The eight Green Triangle sites used in this study were E. The original version of 3-PG was calibrated for E. globulus
globulus plantations planted between 1995 and 1998 and grown in Australia (Sands and Landsberg, 2002; Booth et al.,
described in detail by Benyon and Doody (2004) and 2006; Smettem et al., 2013), and Fontes et al. (2006) adapted
Benyon et al. (2006, 2007, 2008). Mean annual precipi- these parameters for E. globulus grown in Portugal. Almeida
tation varied from 500 to 800 mm, and potential ET was et al. (2004b, 2007) provided a parameterization for E. grandis
980 mm year1 (Wang et al., 2001). grown in Brazil that predicted foliage, stem and root biomass
For all Australian sites, daily maximum and minimum with remarkable accuracy. Esprey et al. (2004) found a very
air temperature, rainfall and solar radiation data were similar set of parameters for E. grandis grown in South Africa.
obtained from the Scientific Information for Land Owners These species-specific parameter sets were used as the basis for
(SILO) Data Drill (Jeffrey et al., 2001) from planting until the applications reported here, with a minor re-calibration for
the last growth measurement at the site or the end of the the new water balance model required because of the interaction
established rotation. SILO data are constructed by spatially between ASW and tree growth. This is described later.
interpolating (on a regular 0·05° grid, with latitude, An important parameter is the specific root volume σ zR,
longitude and elevation as independent variables) daily i.e. the volume of soil explored by unit biomass roots, because
data collected by the Australian Bureau of Meteorology this determines the extent of the soil zone from which water is
(Jeffrey et al., 2001). Daily SILO data were used to taken up by the stand. Data were available for the Brazilian site
calculate monthly average maximum and minimum air for root biomass and root depth as a function of stand age and
temperature and solar radiation, monthly rainfall and for a number of different clones of E. grandis (Almeida, 2003).
number of rain and frost days per month (i.e. minimum These data were used to assign σ zR = 2·5 for E. grandis.
air temperature < 0 °C). Soil texture had not been directly Possible variation in σ zR is discussed later. The same value was
measured in the Green Triangle sites but was taken from used for E. globulus because no alternative data were available.
the Victorian soils database (DNRE, 2002). Soils varied A minor re-calibration for the new water balance model
from sandy loam/clay to clay loam/clay, depth ranged from because of the interaction between ASW and tree growth is
3·9 to 6·0 m, and water table levels varied from 3·3 to 20 m described in the following.

Copyright © 2015 John Wiley & Sons, Ltd. Ecohydrol. 9, 610–630 (2016)
MODEL WATER BALANCE OF FOREST PLANTATIONS IN CONTRASTING ENVIRONMENTS 615

(m day1)
The new input data required to run the model are the

kScond

0·2
0·5
0·5
1·0
0·5
0·5
0·5
1·0
0·5
5·0
0·5
following soil characteristics: soil profile depth (m);
volumetric SWC at saturation, field capacity and wilting
point (m3 m3); saturated hydraulic conductivity (cm h1);
drainage ratio (month1); and parameters ES1 and ES2

(day1)
kDrain

0·10
0·20
0·50
0·20
1·00
0·80
0·80
0·66
0·50
0·50
0·50
characterizing soil evaporation. These are presented in
Table II for each experimental site. The specific root
volume was estimated from root depth and biomass at

Esoil2

0·50
0·60
0·30
0·60
0·60
0·50
0·25
0·30
0·30
0·30
0·30
different ages (Almeida, 2003).
Water balance is dependent on the LAI of a stand, so any
error in the prediction of growth would impact on the

Esoil1

0·20
0·20
0·10
0·10
0·30
0·10
0·10
0·05
0·10
0·10
0·10
performance of the water balance submodel. Accordingly,
because the Aracruz site had a long series of observed LAI,
we tested the model’s performance when predicted LAI

5
5
5
5
8
7
8
8
5
9
5
was replaced with observed LAI in order to reduce
potential errors in the water balance due to errors in the

0·5
0·50
0·50
0·50
0·65
0·60
0·65
0·65
0·50
0·70
0·50

growth predictions. This approach did not produce any

Table II. Values of soil characteristics at each experimental site.


major changes in the results, largely because the original

(m3 m3)
model produced excellent predictions of LAI (Almeida

SWwilt

0·240
0·270
0·380
0·350
0·125
0·235
0·090
0·060
0·290
0·120
0·170
et al., 2004b; Smettem et al., 2013).
Model efficiency (EF) and the root mean square error
(RMSE) were calculated to quantify the model accuracy. For
EF, observed and predicted values of selected variables were

(m3 m3)
SWfcap

0·29
0·32
0·40
0·42
0·26
0·36
0·20
0·09
0·46
0·27
0·30
assessed using Equation (1) (Soares et al., 1995). EF can
vary from 1 (low accuracy) to +1 (high accuracy); when
positive, the model predicts better than average and, when
negative, worse than average. The RMSE [Equation (2)] is a
(m3 m3)
SWsat

0·33
0·34
0·48
0·45
0·32
0·42
0·26
0·39
0·47
0·30
0·32
measure of the precision of the estimates in the same units as
the dependent variable.
 2
∑ni¼1 Y i  Y^ i
0·8
0·8
0·7
0·8
0·8
0·8
0·9
0·9
0·8
0·6
0·6
FR

EF ¼ 1  n  2 (1)
∑i¼1 Y i  Y
SWC measured
Depth at which

sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 2
(m)

2·6
5·7
5·7
2·7
2·7
5·0
1·6
4·5
5·7
8·0
8·0
∑n Y i  Y^ i
RMSE ¼ i¼1
(2)
n

where Yi and Ŷi are the observed and predicted values,


depth (m)
Max root

respectively, Ῡ is the mean of the observed values and n


2·0
4·0
6·0
3·0
3·0
3·0
1·6
4·0
5·7
8·0
5·0

represents the total number of measurements.

Eucalyptus grandis hybrids grown in Brazil


depth (m)

An additional set of biomass observations (at age 7 years) is


Soil

2·6
6·0
6·0
4·3
3·9
5·0
1·8
5·0
6·0

5·0
15·0

now available following Almeida et al. (2004b), and the


original parameters for E. grandis grown in Brazil were
applied with the new model to simulate this extended set of
observations. Root biomass at 7 years was slightly under-
class
Soil

SCL
CL

CL

CL
LS

LS
LS
SL
LS
LS

predicted; the only change required to simulate the extended


data was to reduce root turnover from 0·02 to 0·015
(month1). Simulation of SWC was improved by changing
Scott River
Padthaway

Wellstead

maximum canopy conductance, attained at an LAI of 3·0,


Gummy
Aracruz

from 0·020 to 0·024 (m s1). In Almeida et al. (2004a), the 3-


Bessie

Jasper

Vikki
Jack
DPI
Site

Jill

PG parameter MaxAge in the age-related growth modifier

Copyright © 2015 John Wiley & Sons, Ltd. Ecohydrol. 9, 610–630 (2016)
616 A. C. ALMEIDA AND P. J. SANDS

fage was quoted as 9 years, but fage is in fact not used, and a forest growth from age 1 to 7·5 years. WB2 predicted 12%
large value (>100 years) should be assigned. lower stem mass than did WB1 at the end of the rotation.
Most of the difference occurred from age 5 to 7 years.
Eucalyptus globulus grown in Australia Soil water content was predicted using pseudo-daily and
Few changes to parameters given by Sands and Landsberg daily time steps in WB2 and is compared with observed
(2002) were required to adequately predict the growth of E. data in Figure 3. Prediction accuracy (R2) of SWC was
globulus at the Green Triangle and WA sites: the power in the 86% with a pseudo-daily time step and 90% with a daily
stem mass allometric relationship was changed from 2·4 to time step. WB1 does not predict SWC. The standard error
2·25; the minimum allocation of net primary production (NPP) (SE) of predicted SWC with WB2 was about 14% of the
to roots from 0·25 to 0·1; maximum litterfall rate from 0·027 to maximum ASW at this site (172 mm) with a pseudo-daily
0·015 (month1); and maximum canopy conductance, attained time step and 10% with a daily time step.
and LAI 3·3, from 0·020 to 0·028 m s1 (Schulze et al., 1994, Available soil water was predicted using pseudo-daily and
White et al., 1999); and the influence of fertility on canopy daily time steps in WB2 and also with WB1 and is compared
quantum efficiency was set at 0·6, as used elsewhere. The sites with observed data in Figure 4. Prediction accuracy (R2) of
used by Sands and Landsberg (2002) all had FR = 1. These ASW was 85% with a pseudo-daily time step and 91% with a
changes were based on observed data from these sites and daily time step, but only 67% with WB1 (which uses a
related studies (Benyon et al., 2006; White et al., 2009). monthly time step). The SE of predicted ASW with WB1 was
about 15% of the maximum ASW at this site. For WB2, the
Sensitivity analysis SE of predicted ASW was 11% with a pseudo-daily time step
and 7% with a daily time step.
A sensitivity analysis provides information on how WB2 was particularly effective at predicting soil
sensitive an output is to changes to, or uncertainties in, moisture changes after rainfall events and the reduction
species-specific parameters, site and soil characteristics or of soil moisture by transpiration. In particular, the new
other inputs. We used the concept of relative sensitivity model, using both pseudo-daily and daily time steps,
defined by Brylinsky (1972) and used by Battaglia and predicted the low ASW observed early in the rotation and
Sands (1998) and Esprey et al. (2004). from age 5 years onwards better than WB1 did. This led to
Brylinsky (1972) defined the relative sensitivity λ(X, p) higher water stress and reduced growth at the end of the
of X with respect to p, as the change δX in X produced by a rotation, as was observed. The ability of WB2 to predict
change δp in p relative to the original values of X and p: ASW early in the rotation (Figure 4a) was because root
depth depends on root biomass, which results in a more
p ∂X realistic behaviour of root distribution in the soil profile,
λðX ; pÞ ¼ (3) and hence of water use. The volume of soil occupied by
X ∂p
roots is small after planting but increases as roots develop
Relative sensitivity is zero if X is independent of p and is over time up to the maximum root depth, achieved at age
positive or negative depending on whether an increase in p 3 years. A synthesis of the predicted water balance
results in an increase or a decrease, respectively, in X. components obtained with both WB1 and WB2 using
Large values of λ indicate that X is highly sensitive to p, monthly, pseudo-daily and daily simulations is presented in
while a value of λ close to zero indicates X is essentially Table III for the Aracruz site. This table shows rainfall, ET
independent of p. In particular, if λ = 1, then a 10% change, (i.e. ET, the sum of transpiration, soil evaporation and
for example, in p will lead to a 10% change in X. rainfall interception), drainage and run-off accumulated
Esprey et al. (2004) also provide a simple computational over a period of 7·5 years.
formula for determining λ from data obtained by running Daily estimates of SWC and ASW at the Aracruz site are
the model with the parameter p increased and decreased by presented in Figures 3a and 4a, respectively. The predictions
a small amount δp. We determined λ for a range of model were compared with 206 measurements made in the stand from
outputs with respect to a range of soil characteristics, using age 2 to 7·5 years. Values of ASW were inferred from SWC
δp = 0·1p (i.e. 10% changes in p). measured (at 20-cm intervals with a neutron probe) down to the
root depth estimated by the model. It was assumed that the
minimum observed value of SWC corresponded to the wilting
RESULTS point. The maximum root depth of 2·0 m was reached 3 years
after planting, and the corresponding maximum soil water
Growth and water balance at the Brazilian site holding capacity is 172 mm. The use of the new model with
The model accurately predicted forest growth at the either a pseudo-daily or daily time step improved the estimation
Aracruz site. Figure 2 shows predicted and observed stem of ASW when compared with the original model (which uses a
and root biomass, diameter at breast height (DBH) and LAI monthly time step). It also allows the direct prediction of SWC.

Copyright © 2015 John Wiley & Sons, Ltd. Ecohydrol. 9, 610–630 (2016)
MODEL WATER BALANCE OF FOREST PLANTATIONS IN CONTRASTING ENVIRONMENTS 617

Figure 2. Stand growth predicted with WB2 (–––) and WB1 (– – –) for stem and root mass, leaf area index and DBH at the Aracruz site. Dot points (●)
are observed data.

Figure 3. Comparison of observed and predicted soil water content (SWC) data for the Aracruz site. (a) Observed (●) and time course of SWC predicted
by WB2 with pseudo-daily (­­­­­­­) and daily (–––) time steps and correlation between observed and predicted SWC using WB2 with (b) pseudo-daily
and (c) WB2 daily time steps. The 1-to-1 line is shown (······).

Copyright © 2015 John Wiley & Sons, Ltd. Ecohydrol. 9, 610–630 (2016)
618 A. C. ALMEIDA AND P. J. SANDS

Figure 4. Comparison of observed and predicted available soil water (ASW) data for the Aracruz site. (a) Observed (●) and predicted time course of
ASW by WB1 (------) and WB2 pseudo-daily (······) and daily (–––) time steps and correlation between predicted observed ASW using (b) WB2 pseudo-
daily, (c) WB2 daily and (d) WB1. The 1-to-1 line is shown (······).

Table III. Mean annual water balance components (mm) at the Aracruz site.

Evapotran Soil Canopy Rainfall


Rainfall spiration evaporation transpiration interception Drainage Run-off SWC ASW

Individual components of water balance


WB2 pseudo-daily simulation (mm) 7257 5585 1709 3229 647 1160 441
WB2 daily simulation (mm) 7276 5336 1219 3592 524 1029 790
WB1 monthly simulation (mm) 7257 5224 — — 777 — 1961
Components as % of rainfall
WB2 pseudo-daily simulation (%) 100 77·0 23·5 44·5 8·9 16·0 6·1
WB2 daily simulation (%) 100 73·3 16·8 49·4 7·2 14·2 10·8
WB1 monthly simulation (%) 100 72·0 — — 10·7 — 27·0
Annual averages
Annual average 967 771 710 56
– monthly simulation (mm)
Annual average 967 712 721 63
– daily simulation (mm)

Growth and water balance at the Australian sites Figure 6 compares the observed and predicted ET
(i.e. sum of interception, transpiration and soil evaporation)
The Green Triangle and WA sites were analysed individ- for four sites in the Green Triangle region demonstrating a
ually to evaluate model performance. Both WB1 and WB2 reasonable agreement between observed and predicted except
were applied to the Green Triangle site DPI because it had a for Vikki site. Predictions were made with WB2 using a
long series of observed data. Figure 5 compares nine outputs pseudo-daily time step. Figure 7 compares observed and
from the two models with corresponding observations, and predicted SWC for six sites in the Green Triangle and WA at
this comparison shows WB2 provides a better fit to the which SWC had been observed. Predictions were with WB2
observed data. using a pseudo-daily time step. We compared measured SWC

Copyright © 2015 John Wiley & Sons, Ltd. Ecohydrol. 9, 610–630 (2016)
MODEL WATER BALANCE OF FOREST PLANTATIONS IN CONTRASTING ENVIRONMENTS 619

Figure 5. Temporal variation of observed and predicted stand volume, diameter at breast height, leaf area index, evapotranspiration, soil water content,
canopy interception, available soil water (predicted only), soil evaporation and canopy transpiration for the DPI site in the Green Triangle region.
Observations are ●, predictions by WB1 are (­­­­­­­), and by WB2 are (–––).

with the predictions made by WB2 for similar periods using the soil water balance accumulated over the entire rotation to
data collected at dates close to the estimation periods. each of these new input data. Sensitivity may depend on the
An evaluation of the quality of the predictions of forest stand age at which the analysis is made (e.g. Minunno et al.,
growth, as represented by DBH, is presented in Table IV 2013; Song et al., 2013). For instance, heavy rain at the age at
for the Australian sites. This shows the precision and which the sensitivity analysis is determined can have
accuracy of predicted DBH using the basic statistics of the pronounced effects on transient fluxes of soil water.
observed versus predicted regression lines, model effi- Accordingly, it is generally advisable to base the sensitivity
ciency (EF) and the root mean square residual (RMSE). analysis on variables that are accumulated over the entire
history of the stand, and our results are for variables
Sensitivity analysis accumulated up to the end of the rotation. They clearly
indicate that the SWC at saturation, field capacity and wilting
We ran a sensitivity analysis for the Aracruz site using both
point has strong effects on the water balance and, because this
daily and pseudo-daily time steps to test the sensitivity of the
affects soil water stress, also on stand biomass components.
model to the new site-specific inputs (soil saturation, field
capacity and wilting point, drainage, hydraulic conductivity
and soil evaporation phases 1 and 2). The calculation of
DISCUSSION
sensitivity was based on 10% changes in each input, and
results were very similar for both time steps. Table V shows Water use by Eucalyptus plantations has been a controver-
the sensitivity of the biomass pools and each component of sial issue in various regions (Calder, 1992; Soares and

Copyright © 2015 John Wiley & Sons, Ltd. Ecohydrol. 9, 610–630 (2016)
620 A. C. ALMEIDA AND P. J. SANDS

Figure 6. Comparison of evapotranspiration predicted by WB2 using pseudo-daily time step with observed data for four experimental sites in the Green
Triangle region. The 1-to-1 line is shown (------).

Almeida, 2001; Dye, 1996; Dye and Versfeld, 2007; below which is a root-free zone. There is also an implicit
Dvorak, 2012), particularly aggravated by the expansion surface layer from which soil water is evaporated. Our
of plantations into drier areas. The impacts of this expansion objection is to the use of a true multi-layered soil profile.
on water resources are often unknown, and inappropriate The model does not explicitly consider access to a water
generalizations and extrapolations of the results of other table, e.g. by tap roots. In particular, the simplicity of 3-PG
studies have been made without considering local and does not allow the prediction of the vertical distribution of root
regional biophysical characteristics. In many cases, these biomass. Smettem et al. (2013) suggested increasing the
generalizations are driving policies or imposing restrictions maximum ASW at a site if it is suspected that a stand has access
to expansion of plantations without scientific basis. Soil to groundwater. However, we suggest setting a non-zero
physical characteristics can vary drastically in relatively minimum ASW such that if the root-zone ASW falls below this
small areas and have a strong influence on water availability. value it is assumed the stand draws water from the water table
The lack of detailed maps of soil physical characteristics is so as to maintain water stress above this minimum amount.
obviously a limitation for accurate spatial prediction of Disaggregation of the transpiration flux of WB1 into
plantation water balance (Feikema et al., 2010). canopy transpiration and soil evaporation as presented in
Models such as 3-PG are often used to predict plantation WB2 provides a more realistic estimation of the compo-
growth and water use, and it is imperative that they provide nents of the water balance because they are now separately
reliable predictions of plant and soil water balance. The affected by the environment and are more realistically
modifications (WB2) that we incorporated into the original modelled. These two fluxes can be compared with
3-PG water balance submodel (WB1) were inspired by the measurements such as sap flow from individual trees and
need to improve the water balance predictions. mini-lysimeters used to measure soil evaporation.
The new water balance model assumes a single-layer Based on the results presented by applying WB1 and WB2
homogeneous soil profile, rather than multiple layers as to the same sites, we conclude that the modifications leading
described by Soares and Almeida (2001). However, it does to WB2 improved the estimation of the SWC and ASW in
subdivide the profile into a horizontal zone containing roots, plantations of two species growing in contrasting environ-

Copyright © 2015 John Wiley & Sons, Ltd. Ecohydrol. 9, 610–630 (2016)
MODEL WATER BALANCE OF FOREST PLANTATIONS IN CONTRASTING ENVIRONMENTS 621

Figure 7. Comparison of soil water content predicted by WB2 with a pseudo-daily time step with observed data at six experimental sites in Green
Triangle and Western Australia.

Table IV. Observed versus predicted regression analysis, model efficiency and accuracy for predicted DBH (cm) at the Australian sites.

Bessie DPI Gummy Jack Jasper Padthaway Jill Vikki Wellstead Scott River

Slope 0·65 0·92 0·73 0·61 0·76 1·39 0·71 1·07 1·45 1·14
Intercept 6·22 1·50 4·06 6·14 4·14 4·2 3·87 2·54 6·01 2·41
R2 0·86 0·95 0·98 0·90 0·94 0·98 0·95 0·99 0·99 0·98
N 39 37 67 73 24 17 74 57 7 7
EF 0·78 0·98 0·76 0·74 0·64 0·68 0·80 0·24 0·81 0·74
RMSE 0·50 0·14 1·05 0·79 0·42 0·57 0·64 1·45 1·49 0·98

WB2 with a pseudo-daily time step was used.

Copyright © 2015 John Wiley & Sons, Ltd. Ecohydrol. 9, 610–630 (2016)
622

Table V. Relative sensitivity to various soil characteristics of the stand biomass components (WS, WF and WR), total and available soil water(SWC and ASW) and water balance
fluxes (ET, Transp, Esoil, Rainint, Run-off, Drainage) accumulated over the entire rotation.

Biomass pools Soil water contents Water fluxes

WS WF WR SWC ASW ET Transp Esoil Rainint Run-off Drainage

SWC at saturation 0·27 0·09 0·11 0·15 0·62 0·12 0·23 0·09 0·13 16·5 6·5
SWC at field capacity 0·05 0·03 0·15 0·66 2·9 0·01 0·02 0·02 0·09 7·9 10·5
SWC at wilting point 0·48 0·23 0·38 0·01 4·1 0·29 0·52 0·14 0·24 0·39 1·62

Copyright © 2015 John Wiley & Sons, Ltd.


Rate of drainage (kDrain) 0·03 0·01 0·01 0·02 0·08 0·01 0·02 0·01 0·01 3·7 0·21
Stage 1 soil evaporation (E1) 0·03 0·02 0·02 0·00 0·01 0·00 0·03 0·09 0·02 0·02 0·02
Stage 2 soil evaporation (E2) 0·12 0·12 0·08 0·01 0·03 0·01 0·14 0·34 0·08 0·05 0·05
Soil conductivity (kScond) 0·00 0·00 0·00 0·00 0·01 0·00 0·00 0·00 0·00 0·09 0·05

Data are based on runs with a daily time step. Sensitivities regarded as being exceptionally high are shown bold, while moderately high sensitivities are shown in italic.

Table VI. End-of-rotation predictions for growth, average soil water content and accumulated water balance fluxes for all study sites.

Site Soil evap Trans Rain


A. C. ALMEIDA AND P. J. SANDS

Age Stems WF WR WS Stand volume Average LAI SWC ASW Rainfall ET oration piration Drainage interception Run-off

years tree ha1 t ha1 t ha1 t ha1 m3 ha1 m2 m2 mm monthly average mm accumulated over rotation

Bessie 12 1100 6·8 16·4 91·2 204·1 2·6 1662 60 7966 7689 2290 4057 203 1342 0
DPI 12 950 5·3 14·1 65·0 145·4 1·9 2239 28 6816 6417 2126 3425 300 866 0
Gummy 12 1125 6·5 16·0 97·2 217·6 2·7 1040 51 7605 7658 2159 4152 168 1347 0
Jack 12 1175 6·6 17·7 97·2 232·0 3·0 454 106 7146 7912 2304 4285 120 1323 0
Jasper 12 750 5·5 15·6 75·0 167·8 1·9 1384 83 7218 6918 2588 3439 0 891 0
Padthaway 12 1025 4·1 9·9 52·6 117·7 1·8 180 41 5122 5000 1726 2678 163 596 0
Jill 12 1200 6·8 12·1 87·9 196·7 2·5 298 22 7369 7031 2195 3134 322 1702 0
Vikki 12 925 5·9 17·9 87·6 195·9 2·4 2109 166 7564 7784 2372 4194 248 1218 0
Wellstead 12 1250 5·8 21·5 88·9 198·8 3·0 1349 152 6487 8085 2041 4945 39 1099 0
Scott River 12 1250 5·2 53·3 91·3 170·5 2·4 1347 303 10 695 10 254 2245 6635 62 1374 0
Aracruz 7·5 1111 1·9 17·1 147·8 292·1 2·1 716 60 7257 5585 1709 3228 1160 647 441

Ecohydrol. 9, 610–630 (2016)


MODEL WATER BALANCE OF FOREST PLANTATIONS IN CONTRASTING ENVIRONMENTS 623

ment: there is a better correspondence between the predicted rainfall, drainage from 0% to 16%, soil evaporation from
and observed forest growth and soil water balance. However, 21% to 34% and canopy interception from 9% to 17%. This
we have no explanation as to why the site Vikki has such poor variation in fluxes would be primarily due to the differences
predictions of ET and SWC (Figures 6 and 7) whereas in soil, species characteristics and climate at each site.
predictions of DBH at this site were good (Table IV). As the areas of new plantations increase, there is a clear
This new submodel, however, requires more detailed demand for better understanding of and ability to quantify
information about soil water characteristics and parameters water use by plantations and the effects of water deficit on
than did WB1. The required soil parameters can be derived forest growth. Understanding and quantifying water avail-
from soil texture data with a reasonable degree of confidence ability at the plot and catchment scales allow better
(Clapp and Hornberger, 1978), but we cannot avoid the fact management of wood production by the forests and of the
that more texture-dependent soil data are required. water resources and help avoid potential conflicts caused by
As expected, the use of a daily time step improved the the use of water. The practical implication of the new model
prediction of the water balance (Figures 3 and 4). Our results for forest management appears to be its potential to meet this
are consistent with those of Feikema et al. (2010), who also demand for better understanding and more accurate
tested a modified version of 3-PG using a daily time step. quantification of water used by trees and water production
With the exception of one site (Vikki), our estimations of ET from catchments.
have a similar accuracy to those reported by Feikema et al. The results of the sensitivity analysis (Table V) indicate
(2010). The latter also found that simulations of ASW using that the water balance is highly sensitive to some of the soil
a monthly time step are higher than using a daily time step. parameters required for the new model. Wilting point and
However, our model explicitly takes into account a changing field capacity strongly affect ASW: field capacity and the
root distribution during early stand development. drainage parameter (kD) have respectively negative and
Daily climatic data are not always available. Also, their use positive impacts on the prediction of run-off, and field
for spatial applications could be a limiting factor as increasing capacity also influences drainage and ET. The effects of the
the number of climatic surfaces adds complexity and soil characteristics on tree growth were less pronounced
computational demands. If daily data are not available or if and would be mediated through their effects on ASW, and
its use is inappropriate for other practical reasons, either version hence on the soil water growth modifier.
of the model can be run with a pseudo-daily time step. This The specific root volume σ zR determines the extent of the
treats each month as a series of dry spells of equal length, with soil zone from which water is taken up by the stand. Data
rain falling on the first day of each spell. The use of a pseudo- for root biomass and root depth as a function of stand age
daily time step results in a marked improvement in the ability of and for a number of different clones of E. grandis
the model to predict finer details of the temporal variation of (Almeida, 2003) were used to estimate σ zR, and the
soil water status, even with the original water balance model. resulting values are plotted in Figure 8a for the clone used
Prediction of water balance components for the 11 study in this study. It is clear that σ zR declines with stand age, but
sites shows considerable differences between some sites. there is considerable variation at a given stand age. (Similar
This is shown in Table VI, which also shows an example of values were obtained for other clones but are not shown.) It
the outputs of plantation growth and water balance for the is clear σ zR may vary with stand age, and we would expect
full rotations of the study sites. The only site that produced possible variations with soil texture and between species.
run-off was at Aracruz. Over a full rotation (7·5 years for the We explored the possible effects of changes in σ zR by
Brazilian site and 12 years for the Australian sites), running the model for the Brazilian site using radically
transpiration varied from 41% to 60% of the incoming different values. Figure 8b shows the variation in stem

Figure 8. Analysis of specific root volume. (a) Observed values for Eucalyptus grandis at the Brazilian site as a function of stand age. (b) The effect of
varying specific root volume at this site on the stem biomass (–♦–) and total accumulated ET (–♦–) at the end of the rotation.

Copyright © 2015 John Wiley & Sons, Ltd. Ecohydrol. 9, 610–630 (2016)
624 A. C. ALMEIDA AND P. J. SANDS

biomass and accumulated ET at the end of the rotation with Dick Waring and an anonymous reviewer for their comments
σ zR. It is clear that in this case that values of σ zR >1 are on an earlier draft of the manuscript.
equivalent, but there are strong effects for σ zR < 1. Small
values of σ zR mean the roots are slower in exploring the soil
profile and hence water stress effects are larger later in APPENDIX A
stand growth. In our case, the value used for σ zR was that WATER BALANCE SUBMODEL EQUATIONS
observed for a young stand and hence the predicted root
depth matched observed. Once the roots attain maximum A.1 Hydrological balance
depth, the subsequent value of σ zR is largely irrelevant.
Our objective was to improve the prediction of water The basic equation governing the hydrological balance
balance without losing the simplicity of the 3-PG model; of a soil profile specifies the change in the SWC Θ in some
we demonstrated that the new model achieves this time interval Δt in terms of the rainfall R, possibly
objective. There are advantages and disadvantages in using including any irrigation, and various losses that occur in
the new submodel, and the decision about which submodel that interval. Thus (L&S, Section 7.1),
will be more appropriate should be based on the questions
to be answered, and obviously on the availability of data. Θðt þ Δt Þ ¼ Θðt Þ þ R  I R  E C  ES  qD  qR (A1)
We believe that there is no advantage in using the WB2
submodel if the soil information is very limited or where all units are millimetres (or kilograms per square
inadequate. On the other hand, if a more precise water metre) and the losses are calculated over the time interval
balance is required, this new submodel proves to be a better Δt. The losses are the amount IR of rainfall intercepted by
approach and provides more accurate predictions of the the canopy and subsequently evaporated, canopy transpi-
growth and water use by plantations. Finally, if daily data ration EC, evaporation ES from the soil, the amount qD of
and detailed soils data are both not available, the original water that drains out of the bottom of the profile and
soil water balance model can be improved by using a surface run-off qR. Lateral flow is not considered in 3-PG.
pseudo-daily time step in lieu of a monthly time step. If the soil profile depth is z (m), then the volumetric
SWC θ (m3 m3 or mm m1) of the profile is given by
θ = Θ/z. The relative plant ASW θr is the ratio of the current
CONCLUSIONS plant ASW to the maximum ASW:
Analyses of the predicted growth and water balance in three  
contrasting regions show that the new 3-PG water balance θ  θwp
θr ¼   (A2)
submodel was able to improve the predictions of the water θf c  θwp
balance components at widely contrasting sites, and for two
species. Moreover, use of the pseudo-daily time step still where θwp and θfc are the volumetric SWCs (mm) of the
provides improved simulation of water balance when using profile at wilting point and field capacity, respectively.
the original water balance submodel because it better captures
the dynamics of rainfall distribution. The new submodel
A.2 Soil water stress and the soil water growth modifier
requires more detailed information of the soil characteristics,
and the water balance is sensitive to these parameters. It is our
Water stress is determined by the water content of a zone
intention to make the new model also freely available.
of soil around the roots. The size of this zone changes
dynamically as plants grow and the root system occupies
more of the soil profile. All transpiration is drawn from the
ACKNOWLEDGEMENTS
root zone, which consequently dries out far more rapidly
Richard Benyon and Tanya Doody provided the Green than the bulk soil. Recharge of the root zone is modelled
Triangle data, Don White provided the WA dataset and Fibria (Appendix A.7) by movement of water at a rate
(formerly Aracruz Celulose) supported the data collection proportional to the difference in volumetric SWC in the
from the Aracruz site as published in Almeida (2003). We two zones and depends on the saturated hydraulic
also appreciate Keryn Paul and Anders Siggins for their conductivity of each soil.
contributions to data collation and Phil Polglase for his The soil water growth modifier (fθ) determines the
incentive to write this paper. We thank Joe Landsberg, effects of soil water stress on growth and canopy
Anthony O’Grady and Jacqui England for reviewing the conductance and is based on relative plant ASW (θr) in
manuscript and Sadanandan Nambiar for his comments. We the root zone, not in the entire soil profile. It is defined so
acknowledge the farmers in WA, VIC and SA who allowed us that it varies between zero and one while soil water varies
access to their properties for data collection. We also thank between no plant available water (θr = 0) to unlimited soil

Copyright © 2015 John Wiley & Sons, Ltd. Ecohydrol. 9, 610–630 (2016)
MODEL WATER BALANCE OF FOREST PLANTATIONS IN CONTRASTING ENVIRONMENTS 625

water (θr = 1). Thus, where L is canopy LAI. Because soil evaporation is
separately modelled in WB2, canopy conductance gC
1  ð1  θ r Þn θ declines to zero as L declines to zero and is the product of
f θ ðθ r Þ ¼ (A3)
1 þ ð1  2cθ nθ Þ½ð1  θr Þ=cθ nθ the optimal conductance gCopt (m s1) and 3-PG growth
modifiers, as in L&S [Equation (9.16)], with
where the parameters cθ and nθ are soil texture-dependent
parameters (L&S, Section A2.1).
 
L
A.3 Rainfall interception g Copt ðLÞ ¼ min ; 1 g Cx (A8)
LgCx
The fraction of rainfall intercepted by the canopy is
assumed to be proportional to canopy LAI L, up to some where LgCx is the value of L at which conductance attains its
maximum fraction iRx that occurs at an LAI of LRx (L&S, maximum value gCx (m s1). LgCx and gCx are species-
Section 7.2.2), where LRx and iRx are species-specific specific parameters.
parameters. The rainfall interception loss IR is
A.5 Soil evaporation
I R ¼ iRx minf1; L=LRx gR (A4)
Soil evaporation is modelled using the so-called two-
A.4 Canopy transpiration phase Ritchie (1972) model. Soil evaporation is assumed to
be confined to a shallow upper soil layer that acts as a
Daily canopy transpiration E C (kg m 2 day 1 or barrier to further evaporation as the soil dries. Evaporation
mm day1) is modelled as in the original 3-PG water is high when the soil surface is wet after rain and declines
balance submodel using the Penman–Monteith equation as it dries out. Evaporation from soil in full sun and that
(L&S, Section 7.2.1): from soil shaded by the canopy are treated separately.
In Phase 1 (immediately after a wetting event), the rate
  eS (mm day1) of soil evaporation is the wet-surface rate, e0
gC sφna þ gb ρa cpa D (mm day1). Phase 2 commences once accumulated
EC ¼ h (A5)
λððγ þ sÞgC þ γg aC Þ evaporation ES (mm) since the last wetting event exceeds
an amount ES1, and eS is then assumed to decline
Here, h (s day1) is the day length, φna (W m2) is the daily hyperbolically with increasing evaporation accumulated
average net absorbed radiation, D (Pa) is the atmospheric above that in Phase 1. The rate of decline is described by a
VPD above the canopy, gC (m s1) is the canopy conductance parameter ES2, which is the amount of evaporation in Phase
and gaC (m s1) is the aerodynamic boundary layer 2 required to reduce eS by 50%. Thus,
conductance of the canopy. Other quantities are the latent
heat λ = 2·45 MJ kg1 of vaporization of water, the density 8
ES ≤ ES1
dE S <
e0
ρa = 1·204 kg m3 and specific heat cpa = 1004 J kg1 K1 of eS ¼ ¼ e0 (A9)
dry air, the psychrometric constant γ = 66·1 Pa K1 and the dt : E S1 < E S
1 þ ðE S  E S1 Þ=E S2
derivative s = 145 Pa K1 (at 20 °C) of the saturated vapour
pressure with respect to temperature.
Because EC is canopy transpiration and does not include which can be integrated to give
evaporation from the soil, φna is the radiation absorbed by
the canopy, not incident upon it. Almeida et al. (2003) (
found e0 t t ≤ t S1
ES ðt Þ ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 
ES1 þ E S2 1 þ 2ðe0 =E S2 Þðt  t S1 Þ  1 t > t S1
φna ¼ a þ bφi (A6)
(A10)
2
where φi (W m ) is the average daily intercepted radiation,
and a and b have values close to 0 and 0·8, respectively. where tS1 = ES1/e0 is the duration of Phase 1 evaporation
From Beers law, φi is related to daily insolation Q (L&S, Section 7.2.4). The wet-surface rate e0 is calculated
(MJ m2 day1) by using the Penman–Monteith equation [Equation (A5)] and
is driven by radiation incident on the soil and the VPD
above the soil, taking into account the presence of a canopy
Q  and that prior to canopy closure part of the soil is in direct
φi ¼ 106 1  ekL ; (A7)
h sun. The analogue of gC is very large, but aerodynamic

Copyright © 2015 John Wiley & Sons, Ltd. Ecohydrol. 9, 610–630 (2016)
626 A. C. ALMEIDA AND P. J. SANDS

conductance gas above the soil is smaller than canopy zone and a root zone. The root zone is that volume of soil
aerodynamic conductance, e.g. gas ≈ 0·01gac. There are from which the current stock of roots can access soil water.
three new parameters: ES0 and ES1, which we assume are Its depth zR (m) is determined from the root biomass WR
texture-dependent soil properties, and the aerodynamic (t ha1) by
conductance gAs.
The water balance submodel keeps track of the
accumulated net evaporation ES from the soil, as this zR ¼ minf0:1σ zR W R =ð1  pstones Þ; zRx g (A12)
determines subsequent soil evaporation. Because of this,
there is no need to explicitly include a separate surface soil
layer in the model. When a rainfall event occurs, ES is where σ zR (m3 kg1) is the volume of soil explored by 1 kg of
reduced by the amount of throughfall (subject to ES ≥ 0). root biomass and the 0·1 converts tons per hectare to
This forms the initial condition ES0 for the subsequent kilograms per square metre. As roots grow, the root-zone
period of soil evaporation, and Equation (A10) is inverted volume increases at the expense of the root-free volume. The
and solved for an equivalent initial time t0. If the time to the SWCs of these two volumes are adjusted to take this into
next rainfall event is tR, soil evaporation in the intervening account.
period is ES(tR + t0)  ES0. For the daily time step model, Water also moves from the root-free zone into the root
tR = 1. For monthly or pseudo-daily time steps, tR is days in zone along the water pressure gradient induced by
the month/number of rain days, and soil evaporation is differences in their volumetric SWCs. If two regions have
computed separately for each rainfall event and then volumetric SWCs θ1 and θ2, with θ1 < θ2, the flux of water
summed over all events. from regions 2 to 1 is assumed to be
Although the Ritchie model was originally empirical, it
can be derived from physical principles after making some F V ¼ k S ðθ1  θ2 Þ (A13)
simple approximations. The parameters ES0 and ES1 are
then expressed explicitly in terms of various textural
properties of the soil. where FV (m3water m2 day1) is the volume of water
crossing unit area in a day and kS (m day1) is a texture-
A.6 Run-off and drainage dependent measure of the hydraulic conductivity of the
soil. Root-zone recharge is modelled by assuming that
If rainfall saturates the soil, i.e. volumetric water during recharge the volumes of the two zones are constant,
content θ exceeds saturation (θsat), the excess above the total amount of water in the two zones is constant and
saturation is deemed to be surface run-off. The remaining they share a common area A (m2). Application of Equation
soil water is available for transpiration and evaporation. (A13) then shows that the temporal variation of the SWC
If θ exceeds field capacity (θfc), this excess is available Θrz(t) of the root zone is given by
for drainage, which proceeds at a rate proportional to the
excess, where the rate constant kDrain (day1) is a texture-
dependent soil property. Drainage qD over t days is then
Θrz0 V nr  Θnr0 V rz t=tS0
given by Θrz ðt Þ ¼ e
V rz þ V nr
(A14)
   V rz
qD ðt Þ ¼ Θ0  V Soil θf c 1  ek Drain t (A11) þ ðΘrz0 þ Θnr0 Þ
V rz þ V nr

where Θ0 is the initial SWC, VSoil is the volume of soil


and θfc is the volumetric soil content at field capacity. Here, Θrz0 and Θnr0 are the SWCs of the root and non-
Drainage is first determined from the root zone into the root zones at time t = 0, Vrz and Vnr are the volumes of the
root-free zone and then from the root-free zone out of root and non-root zones and tS0 is the time constant for
the bottom of the profile. For a daily time step, t = 1; for water movement and is given by
the pseudo-daily time step, t = number of days between
rainfall events, i.e. days in month/rain days; and with a
monthly time step, t = number of days in the month.
t S0 ¼ V rz V nr =k S AðV rz þ V nr Þ (A15)
A.7 Root-zone growth and recharge

The root-zone concept introduced in this paper is simple Equations (A14) and (A15) determine the SWC of the
and pragmatic. The soil profile is divided into a root-free root zone after t days.

Copyright © 2015 John Wiley & Sons, Ltd. Ecohydrol. 9, 610–630 (2016)
MODEL WATER BALANCE OF FOREST PLANTATIONS IN CONTRASTING ENVIRONMENTS 627

APPENDIX B
3-PG PARAMETERS FOR E. GLOBULUS AND E. GRANDIS HYBRIDS APPLIED IN THIS STUDY
Meaning/comments Name Units E. grandis E. globulus

Biomass partitioning and turnover


Allometric relationships and partitioning
Foliage : stem partitioning ratio at D = 2 cm pFS2 — 0·7 1·0
Foliage : stem partitioning ratio at D = 20 cm pFS20 — 0·100 0·150
Constant in the stem mass-versus-diameter relationship aWS — 0·049 0·095
Power in the stem mass-versus-diameter relationship nWS — 2·822 2·25
Constant in the stem mass-versus-diameter relationship aWR — 0 0
Power in the stem mass-versus-diameter relationship nWR — 0 0
Maximum fraction of NPP to roots pRx — 0·6 0·8
Minimum fraction of NPP to roots pRn — 0·07 0·20
Fraction of total root allocation going to fine roots pRfine — 0·01 0·5
Volume of soil accessed by 1 kg of root dry matter spRootVol m3 kg1 root 2·5 2.5
Litterfall and root turnover
Maximum litterfall rate gammaF1 month1 0·12 0·015
Litterfall rate at t = 0 gammaF0 month1 0·00169 0·00100
Age at which litterfall rate has median value tgammaF months 13 12
Average monthly branch and bark litterfall rate gammaS month1 0 0
Average monthly coarse root turnover rate gammaR month1 0·02 0·015
Average monthly fine root turnover rate gammaRf month1 0·25 0·25
NPP and conductance modifiers
Temperature modifier (gmTemp)
Minimum temperature for growth Tmin °C 8·0 8·5
Optimum temperature for growth Topt °C 25·0 16·0
Maximum temperature for growth Tmax °C 36·0 40·0
Frost modifier (gmFrost)
Days production lost per frost day kF days 0 0
Fertility effects (gmNutr)
Value of m when FR = 0 m0 — 0 0
Value of fNutr when FR = 0 fN0 — 0·6 0·6
Power of (1  FR) in gmNutr fNn — 1 1
Atmospheric CO2 effects
Ratio of alpha at 700 and 350 ppm gmCalpha700 — 1·4 1·4
Ratio of canopy conductance at 700 and 350 ppm gmCg700 — 0·7 0·7
Salinity effects (gmSalt)
Salinity below which no effects of salt on growth EC0 dS m1 999 999
Salinity above which growth ceases EC1 dS m1 999 999
Power of EC in gmSalt ECn — 1 1
Age modifier (gmAge)
Maximum stand age used in age modifier MaxAge years 20 50
Power of relative age in function for fAge nAge — 4 4
Relative age at fAge = 0·5 rAge — 0·95 0·95
Stem mortality and self-thinning
Mortality rate for large t gammaN1 % year1 0 0
Seedling mortality rate (t = 0) gammaN0 % year1 0 0
Age at which mortality rate has median value tgammaN years 0 0
Shape of mortality response ngammaN — 1 1
Maximum stem mass per tree at 1000 trees per hectare wSx1000 kg tree1 300 300
Power in self-thinning rule thinPower — 1·5 1·5
Fraction mean single-tree foliage biomass lost per dead tree mF — 0 0
Fraction mean single-tree root biomass lost per dead tree mR — 0·2 0·2
Fraction mean single-tree stem biomass lost per dead tree mS — 0·5 0·2
Canopy structure and processes
Specific leaf area
Specific leaf area at age 0 SLA0 m2 kg1 11 11
Specific leaf area for mature leaves SLA1 m2 kg1 8 4
Age at which specific leaf area = (SLA0 + SLA1)/2 tSLA years 2·5 2·5
Light interception and VPD attenuation
Extinction coefficient for absorption of PAR by canopy k — 0·5 0·5
Age at canopy cover fullCanAge years 3 3
(Continues)

Copyright © 2015 John Wiley & Sons, Ltd. Ecohydrol. 9, 610–630 (2016)
628 A. C. ALMEIDA AND P. J. SANDS

APPENDIX B (Continued)
Meaning/comments Name Units E. grandis E. globulus

LAI for 50% reduction of VPD in canopy cVPD0 5 5


Rainfall interception
Maximum proportion of rainfall evaporated from canopy MaxIntcptn — 0·15 0·20
LAI for maximum rainfall interception LAImaxIntcptn — 3 3
Production and respiration
Canopy quantum efficiency alphaCx molC/molPAR 0·068 0·060
Edge tree growth % enhancement edgeEffect — 0 0
Ratio NPP/gross primary production Y — 0·47 0·47
Conductance
Maximum stomatal conductance gSx m s1 0·008 0·008
Radiation for gS = gSx/2 IgS W m2 100 100
LAI for maximum canopy conductance LAIgcx — 3·00 3·33
Minimum canopy conductance MinCond m s1 0 0
Maximum canopy conductance MaxCond m s1 0·024 0·026
Defines stomatal response to VPD CoeffCond mbar1 0·045 0·050
Canopy aerodynamic conductance gAc m s1 0·2 0·2
Soil aerodynamic conductance gAs m s1 0·01 0·02
Wood and stand properties
Branch and bark fraction (fracBB)
Branch and bark fraction at age 0 fracBB0 — 0 0·45
Branch and bark fraction for mature stands fracBB1 — 0 0·15
Age at which fracBB = (fracBB0 + fracBB1)/2 tBB years 0 2
Basic density
Minimum basic density – for young trees rho0 t m3 0·48 0·40
Maximum basic density – for older trees rho1 t m3 0·52 0·45
Age at which rho = (rhoMin + rhoMax)/2 tRho years 5 3
Stem height
Constant in the stem height relationship aH — 0·42 0·39
Power of DBH in the stem height relationship nHB — 1·50 1·41
Power of stocking in the stem height relationship nHN — 0 0·00001
Stem volume
Constant in the stem volume relationship aV — 0 0
Power of DBH in the stem volume relationship nVB — 0 0
Power of stocking in the stem volume relationship nVN — 0 0
Conversion factors
Intercept of net-versus-solar radiation relationship Qa W m2 8·85 0
Slope of net-versus-solar radiation relationship Qb — 0·82 0·8
Molecular weight of dry matter gDM_mol gDM mol1 24 24
Conversion of solar radiation to PAR molPAR_MJ mol MJ1 2·3 2·3

REFERENCES a rotation for pulp production. Forest Ecology and Management 251:
10–21. DOI:10.1016/j.foreco.2007.06.009.
Almeida AC. 2003. Application of a process-based model for predicting Almeida AC, Sands PJ, Bruce J, Siggins AW, Leriche A, Battaglia M,
an explaining growth in Eucalyptus plantations, PhD Thesis, The Batista TR. 2009. Use of a spatial process-based model to quantify
Australian National University, 232 pp. forest plantation productivity and water use efficiency under climate
Almeida AC, Landsberg JJ. 2003. Evaluating methods of estimating change scenarios. 18th World IMACS / MODSIM Congress, Cairns,
global radiation and vapor pressure deficit using a dense network of 1816–1822.
automatic weather stations in coastal Brazil. Agricultural and Forest Almeida AC, Siggins A, Batista TR, Beadle C, Fonseca S, Loos R. 2010.
Meteorology 118: 237–250. DOI:10.1016/S0168-1923(03)00122-9. Mapping the effect of spatial and temporal variation in climate and soils
Almeida AC, Landsberg JJ, Sands PJ, Ambrogi MS, Fonseca S, Barddal on Eucalyptus plantation production with 3-PG, a process-based growth
SM, Bertolucci FL. 2004a. Needs and opportunities for using a process- model. Forest Ecology and Management 259: 1730–1740.
based productivity model as a practical tool in Eucalyptus plantations. DOI:10.1016/j.foreco.2009.10.008.
Forest Ecology and Management 193: 167–177. DOI:10.1016/j. Battaglia M, Sands P. 1998. Application of sensitivity analysis to a model
foreco.2004.01.044. of Eucalyptus globulus plantation productivity. Ecological Modelling
Almeida AC, Landsberg JJ, Sands PJ. 2004b. Parameterisation of 3-PG 111: 237–259. DOI:10.1016/S0304-3800(98)00114-8.
model for fast-growing Eucalyptus grandis plantations. Forest Ecology Battaglia M, Sands P, White D, Mummery D. 2004. CABALA: a linked
and Management 193: 179–195. DOI:10.1016/j.foreco.2004.01.029. carbon, water and nitrogen model of forest growth for silvicultural
Almeida AC, Soares JV, Landsberg JJ, Rezende GD. 2007. Growth and decision support. Forest Ecology and Management 193: 251–282.
water balance of Eucalyptus grandis hybrid plantations in Brazil during DOI:10.1016/j.foreco.2004.01.033.

Copyright © 2015 John Wiley & Sons, Ltd. Ecohydrol. 9, 610–630 (2016)
MODEL WATER BALANCE OF FOREST PLANTATIONS IN CONTRASTING ENVIRONMENTS 629

Benyon RG, Doody T. 2004. Water use by tree plantations in south east Landsberg JJ, Waring RH, Coops NC. 2003. Performance of the forest
south Australia. Technical Report No. 148, 1–26, Australia, CSIRO. productivity model 3-PG applied to a wide range of forest types. Forest
Benyon R, Theiveyanathan S, Doody TM. 2006. Impacts of tree Ecology and Management 172: 199–214. DOI:10.1016/S0378-1127
plantations on groundwater in south-eastern Australia. Australian (01)00804-0.
Journal of Botany 54: 181–192. DOI:10.1071/BT05046. Landsberg JJ. 2003. Modelling forest ecosystems: state-of-the-art,
Benyon R, England J, Eastham J, Polglase P, White D. 2007. Tree water challenges and future directions. Canadian Journal of Forest Research
use in forestry compared to other dry-land agricultural crops in the 33: 385–397. DOI:10.1139/X02-129.
Victorian context. Report prepared for the Department of Primary Landsberg JJ, Sands PJ. 2010. Physiological Ecology of Forest Production.
Industries Victoria to promote scientific knowledge in this area. Ensis Principles, Processes and Models. Elsevier: London, 352 pp.
Technical Report No. 159, 50 p. Mendham DS, White DA, Battaglia M, McGrath JF, Short TM, Ogden
Benyon RG, Doody TM, Theiveyanathan S, Koul V. 2008. Plantation forest GN, Kinal J. 2011. Soil water depletion and replenishment during first-
water use in Southwest Victoria. Technical Report No. 164, CSIRO, 109 pp. and early second-rotation Eucalyptus globulus plantations with deep
Booth TH., Paul KI, Jovanovic T. 2006. Predicting potential plantation soil profiles. Agricultural and Forest Meteorology 151: 1568–1579.
growth in the South West Goulburn Broken. Final Report May 2006, DOI:10.1016/j.agrformet.2011.06.014.
Client Report No. 1669. Miehle P, Battaglia M, Sands PJ, Forrester DI, Feikema PM, Livesley SJ,
Brylinsky M. 1972. Steady-state sensitivity analysis of energy flow in a Morris JD, Arndt S. 2009. A comparison of four process-based model
marine ecosystem. In Systems Analysis and Simulation in Ecology, vol. and statistical regression model to predict growth of Eucalyptus
2, Patten BC (ed). Academic Press: New York; 81–101. globulus plantations. Ecological Modelling 220: 734–746.
Calder IR. 1992. Water use of Eucalypts – a review. In Growth and Water Use DOI:10.1016/j.ecolmodel.2008.12.010.
of Forest Plantations, Calder IR, Hall RL, Adlard PG (eds). John Wiley & Mielke MS, Oliva MA, Barros NF, Penchel RM, Martinez CA, Almeida
Sons Ltd: Chichester, New York, Brisbane, Toronto, Singapore; 167–179. AC. 1999. Stomatal control of transpiration in the canopy of a clonal
Clapp RB, Hornberger GM. 1978. Empirical equations for some soil Eucalyptus grandis plantation. Trees 13: 152–160.
hydraulic properties. Water Resources Research 14: 601–604. Minunno F, Van Oijen M, Cameron DR, Cerasoli S, Pereira JS, Tomé M.
DOI:10.1029/WR014i004p00601. 2013. Using a Bayesian framework and global sensitivity analysis to
Coops NC, Waring RH. 2001. The use of multiscale remote sensing identify strengths and weaknesses of two process-based models
imagery to derive regional estimates of forest growth capacity using 3- differing in representation of autotrophic respiration. Environmental
PGS. Remote Sensing and Environment 75: 324–334. DOI:10.1016/ Modelling & Software 42: 99–115. DOI:10.1016/j.envsoft.2012.12.010.
S0034-4257(00)00176-0. Paul KI, Polglase PJ, Snowdon P, Theiveyanathan S, Raison J, Grove T,
Coops NC. 1999. Improvement in predicting stand growth of Pinus Rance S. 2006. Calibration and uncertainty analysis of a carbon
radiata (D. Don) across landscapes using NOAA AVHRR and Landsat accounting model to stem wood density and partitioning of biomass for
MSS imagery combined with a forest growth process model (3-PGS). Eucalyptus globulus and Pinus radiata. New Forests 31: 513–533.
Photogrammetric Engineering & Remote Sensing 65: 1149–1156. DOI:10.1007/s11056-005-2740-4.
DNRE. 2002. The 1:125,000 statewide soil attribute coverage documen- Paul KI, Booth TH, Jovanovic T, Sands PJ, Morris JD. 2007. Calibration
tation. Version 1.1, June 2002, Centre for Land Protection Research, of the forest growth model 3-PG to eucalypt plantations growing in low
Department of Natural Resources and Environment, Victoria. rainfall regions of Australia. Forest Ecology and Management 243:
Dvorak WS. 2012. Water use in plantations of eucalypts and pines: a 237–247. DOI:10.1016/j.foreco.2007.03.029.
discussion paper from a tree breeding perspective. International Pérez-Cruzado C, Muñoz-Sáez F, Basurco F, Riesco G, Rodríguez-Soalleiro
Forestry Review 14: 110–119. DOI:10.1505/146554812799973118. R. 2011. Combining empirical models and the process-based model 3-PG
Dye PJ. 1996. Response of Eucalyptus grandis trees to soil water deficits. to predict Eucalyptus nitens plantations growth in Spain. Forest Ecology
Tree Physiology 16: 233–238. DOI:10.1093/treephys/16.1-2.233. and Management 262: 1067–1077. DOI:10.1016/j.foreco.2011.05.045.
Dye PJ. 2001. Modelling growth and water use in four Pinus patula stands Ritchie JT. 1972. Model for predicting evaporation from row crop with
with the 3-PG model. Southern African Forestry Journal 191: 53–63. incomplete cover. Water Resources Research 8: 1204–1213.
Dye P, Versfeld D. 2007. Managing the hydrological impacts of South Rodríguez R, Espionosa M, Real P, Inzunza J. 2002. Analysis of productivity
African plantation forests: An overview. Forest Ecology and Manage- of radiata pine plantations under different silvicultural regimes using 3-PG
ment 251: 121–128. DOI:10.1016/j.foreco.2007.06.013. process-based model. Australian Forestry 65: 165–172.
Esprey LJ, Sands PJ, Smith CW. 2004. Understanding 3-PG using a Rodriguez R, Real P, Espinosa M, Perry DA. 2009. A process-based
sensitivity analysis. Forest Ecology and Management 193: 235–250. model to evaluate site quality for Eucalyptus nitens in the Bio-Bio
DOI:10.1016/j.foreco.2004.01.032. Region of Chile. Forestry 82: 149–162. DOI:10.1093/forestry/cpn045.
Feikema PM, Morris JD, Beverly CR, Collopy JJ, Baker TG, Lane PNJ. Running SW, Coughlan JC. 1988. A general model of forest ecosystem
2010. Validation of plantation transpiration in south-eastern Australia processes for regional applications I. Hydrologic balance, canopy gas
estimated using the 3PG+ forest growth model. Forest Ecology and exchange and primary production processes. Ecological Modelling 42:
Management 260: 663–678. DOI:10.1016/j.foreco.2010.05.022. 125–154. DOI:10.1016/0304-3800(88)90112-3.
Fontes L, Landsberg J, Tomé J, Tomé M, Pacheco CA, Soares P, Araujo Sands PJ. 2004. 3PGpjs vsn 2.4 – a user-friendly interface to 3-PG, the
C. 2006. Calibration and testing of a generalized process-based model Landsberg and Waring model of forest productivity. Technical Report.
for use in Portuguese eucalyptus plantations. Canadian Journal of No. 140, CRC Sustainable Production Forestry, Hobart.
Forestry Research 36: 3209–3221. DOI:10.1139/x06-186. Sands PJ, Landsberg JJ. 2002. Parameterisation of 3-PG for plantation
Forrester DI, Tang X. 2015. Analysing the spatial and temporal dynamics grown Eucalyptus globulus. Forest Ecology and Management 163:
of species interactions in mixed-species forests and the effects of stand 273–292. DOI:10.1016/S0378-1127(01)00586-2.
density using the 3-PG model. Ecological modelling. DOI:10.1016/j. Saxton KE, Rawls WJ. 2006. Soil water characteristic estimates by texture
ecolmodel.2015.07.010. and organic matter for hydrologic solutions. Soil Science Society of
Gonzalez-Garcia M, Almeida AC, Hevia A, Majada J, Beadle C. 2015. America 70: 1569–1578.
Application of a process-based model for predicting the productivity of Schulze ED, Kelliher FM, Korner C, Lloyd J, Leuning R. 1994. Relationships
Eucalyptus nitens plantations grown for bioenergy in Spain. Global among maximum stomatal conductance, ecosystem surface conductance,
Change Biology - Bioenergy. DOI:10.1111/gcbb.12256. carbon assimilation rate, and plant nitrogen nutrition: a global ecology
Jeffrey SJ, Carter JO, Moodie KB, Beswick AR. 2001. Using spatial scaling exercise. Annual Review of Ecological Systems 25: 629–660.
interpolation to construct a comprehensive archive of Australian climate Smettem KRJ, Waring RH, Callow N, Wilson M, Mu Q. 2013. Satellite-
data. Environmental Modelling & Software 16: 309–330. DOI:10.1016/ derived estimates of forest leaf area index in South-west Western
S1364-8152(01)00008-1. Australia are not tightly coupled to inter-annual variations in rainfall:
Landsberg JJ, Waring RH. 1997. A generalised model of forest implications for groundwater decline in a drying climate. Global
productivity using simplified concepts of radiation-use efficiency, Change Biology 19: 2401–2412. DOI:10.1111/gcb.12223.
carbon balance and partitioning. Forest Ecology and Management 95: Smethurst PJ, Almeida AC, Loos RA. 2015. Stream flow unaffected by
209–228. DOI:10.1016/S0378-1127(97)00026-1. Eucalyptus plantation harvesting implicates water use by native forest

Copyright © 2015 John Wiley & Sons, Ltd. Ecohydrol. 9, 610–630 (2016)
630 A. C. ALMEIDA AND P. J. SANDS

streamside reserve. Journal of Hydrology Regional Studies 3: 187–198. Wang QJ, Chiew FHS, McConachy FLN, James R, de Hoedt GC, Wright
DOI:10.1016/j.ejrh.2014.11.002. WJ. 2001. Climatic Atlas of Australia: Evapotranspiration. Bureau of
Soares JV, Almeida AC. 2001. Modeling the water balance and soil water Meteorology: Australia.
fluxes in a fast growing Eucalyptus plantation in Brazil. Journal of White DA, Beadle CL, Sands PJ, Worledge D, Honeysett JL. 1999. Quantifying
Hydrology 253: 130–147. DOI:10.1016/S0022-1694(01)00477-2. the effect of cumulative water stress on stomatal conductance of Eucalyptus
Soares P, Tomé M, Skovsgaard JP, Vanclay JK. 1995. Evaluating a growth globulus and Eucalyptus nitens: a phenomenological approach. Australian
model for forest management using continuous forest inventory data. Journal of Plant Physiology 26: 17–27. DOI:10.1071/PP98023.
Forest Ecology and Management 71: 251–265. DOI:10.1016/0378-1127 White DA, Crombie DS, Kinal J, Battaglia M, McGrath JF, Mendham DS,
(94)06105-R. Walker SN. 2009. Managing productivity and drought risk in Eucalyptus
Song X, Bryan BA, Almeida AC, Paul KI, Zhao G, Ren Y. 2013. Time- globulus plantations in south-western Australia. Forest Ecology and
dependent sensitivity of a process-based ecological model. Ecological Management 259: 33–44. DOI:10.1016/j.foreco.2009.09.039.
Modelling 265: 114–123. DOI:10.1016/j.ecolmodel.2013.06.013. Whitehead D, Beadle C. 2004. Physiological regulation of productivity
Stape JL, Ryan MG, Binkley D. 2004. Testing the utility of the 3-PG and water use in Eucalyptus: a review. Forest Ecology and
model for growth of Eucalyptus grandis x urophylla with natural and Management 193: 113–140. DOI:10.1016/j.foreco.2004.01.026.
manipulated supplies of water and nutrients. Forest Ecology and Williams J, Prebble RE, Williams WT, Hignett CT. 1983. The influence of
Management 193: 219–234. DOI:10.1016/j.foreco.2004.01.031. texture, structure, and clay mineralogy on the soil moisture characteristic.
Tickle PK, Coops NC, Hafner SD, Team TBS. 2001. Assessing forest Australian Journal of Soil Research 21: 15–32. DOI:10.1071/SR9830015.
productivity at local scales across a native eucalypt forest using a Winjum JK, Schroeder PE. 1997. Forest plantations of the world: their
process model, 3PG-SPATIAL. Forest Ecology and Management 152: extent, ecological attributes, and carbon storage. Agricultural and
275–291. DOI:10.1016/S0378-1127(00)00609-5. Forest Meteorology 84: 153–167.

Copyright © 2015 John Wiley & Sons, Ltd. Ecohydrol. 9, 610–630 (2016)
Copyright of Ecohydrology is the property of Wiley-Blackwell and its content may not be
copied or emailed to multiple sites or posted to a listserv without the copyright holder's
express written permission. However, users may print, download, or email articles for
individual use.

You might also like