You are on page 1of 11

Advances in Water Resources 74 (2014) 91–101

Contents lists available at ScienceDirect

Advances in Water Resources


journal homepage: www.elsevier.com/locate/advwatres

Root water compensation sustains transpiration rates in an Australian


woodland
Parikshit Verma a,b, Steven P. Loheide II c, Derek Eamus d,b, Edoardo Daly a,b,⇑
a
Department of Civil Engineering, Monash University, Clayton, Australia
b
National Centre for Groundwater Research and Training, Flinders University, Adelaide, Australia
c
Department of Civil Engineering, University of Wisconsin-Madison, WI, USA
d
Climate Change Cluster and School of the Environment, University of Technology, Sydney, Australia

a r t i c l e i n f o a b s t r a c t

Article history: We apply a model of root-water uptake to a woodland in Australia to examine the regulation of transpi-
Received 6 April 2013 ration by root water compensation (i.e., the ability of roots to regulate root water uptake from different
Received in revised form 19 August 2014 parts of the soil profile depending on local moisture availability). We model soil water movement using
Accepted 19 August 2014
the Richards equation and water flow in the xylem with Darcy’s equation. These two equations are cou-
Available online 28 August 2014
pled by a term that governs the exchange of water between soil and root xylem as a function of the dif-
ference in water potential between the two. The model is able to reproduce measured diurnal patterns of
Keywords:
sap flux and results in leaf water potentials that are consistent with field observations. The model shows
Root water compensation
Transpiration
that root water compensation is a key process to allow for sustained rates of transpiration across several
Hydrological modeling months. Scenarios with different root depths showed the importance of having a root system deeper than
Eco-hydrology about 2 m to achieve the measured transpiration rates without reducing the leaf water potential to levels
inconsistent with field measurements. The model suggests that the presence of more than 5% of the root
system below 0:6 m allows trees to maintain sustained transpiration rates keeping leaf water potential
levels within the range observed in the field. According to the model, a large contribution to transpiration
in dry periods was provided by the roots below 0:3 m, even though the percentage of roots at these
depths was less than 40% in all scenarios.
Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction water content [10]. These passive mechanisms of root water


uptake and release are key drivers in regulating water use by veg-
Water taken up by plant roots for transpiration constitutes a etation, becoming especially important in shallow groundwater
significant portion of the hydrological cycle, largely determining environments [11] or duplex soils, where deeper soil layers expe-
the exchange of water, carbon and energy between the land rience large soil water potential (i.e., near saturation) when com-
surface and the atmosphere [1]. The accurate prediction of root pared to the soil near the surface [12].
water uptake is thus important in hydrological and climatological These two mechanisms have been discussed separately in the
applications. literature, with hydraulic redistribution being associated with an
Mechanisms associated with root water uptake that have been imbalance in water potentials across the root system, and root
observed in the field and have received recently renewed interest water compensation being interpreted as a defensive mechanism
are hydraulic redistribution and root water compensation [2–9]. against water stress [13]. Recently, the definition of these two
Hydraulic redistribution refers to the movement of water from soil mechanisms has been merged, by acknowledging their common
layers with higher soil water potential to those with lower soil driving force, i.e., the non uniform water potential distribution
water potential through the root system, while root water com- across the root system [14,15]. Despite their possible different def-
pensation refers to the ability of plants to adjust their distribution initions, it is now recognized that the inclusion of these mecha-
of uptake of water through the soil profile as a function of local soil nisms in mathematical models is recommended to accurately
describe root water uptake [15].
Most of the available models commonly used in hydrological
⇑ Corresponding author at: Department of Civil Engineering, Monash University, applications describe root water uptake as a sink term in the
Clayton, Australia. Richards equation for soil–water flow, and the rate of water
E-mail address: edoardo.daly@monash.edu (E. Daly).

http://dx.doi.org/10.1016/j.advwatres.2014.08.013
0309-1708/Ó 2014 Elsevier Ltd. All rights reserved.
92 P. Verma et al. / Advances in Water Resources 74 (2014) 91–101

  q 
extraction at different depths depends on the amount of water jzj z
rðzÞ ¼ 1 exp jzj  d 6 z 6 0; ð3Þ
available in the soil, the fraction of total roots present in different d d
soil layers and the potential rate at which plant can extract water
where qz ½ is an empirical parameter expressing the decrease of
under unstressed conditions [16–18]. Root water compensation
the root mass with depth.
has been embedded in the sink term by using formulations to
The rate of root water uptake reduces in low soil moisture con-
increase water uptake from wetter parts of the root system to com-
ditions. Additionally, when soil becomes dry, an air gap forms
pensate lower water uptake from roots in drier parts of the soil
between soil and roots, thereby decreasing conductivity between
[13,18–21]. Other models relate root water uptake to root water
the two. To model this reduction in soil-root conductance, we
potential; some assume a defined root water potential distribution
use the function [37]:
across the soil [10,22–25] and others also model water flow
8
through the root xylem [26–29]. Root water uptake is commonly
<0
> h 6 h1 ;
assumed to depend on the difference between soil and root water f ðhÞ ¼ ðh  h1 Þ=ðh2  h1 Þ h1 < h 6 h2 ; ð4Þ
potential. Other formulations are available. For example, van Lier >
:
1 h > h2 ;
et al. [30] expressed water uptake as a function of the difference
between the flux matric potential in the soil and the rhizosphere. where h1 is the volumetric soil moisture content below which root
This approach has been also used to give a mechanistic interpreta- water uptake ceases, and h2 represents the volumetric soil moisture
tion of some of the earlier, empirical definitions of root water com- content below which root water uptake starts decreasing. We did
pensation [31]. More sophisticated models also consider root [32] not consider any reduction in root water uptake due to oxygen
or whole plant hydraulic architecture [14,33–35]. Although these stress.
models permit a more detailed description of the soil water In relation to vegetation, we consider above- and below-ground
dynamics, they are computationally demanding and require a large xylem as a porous medium and thus describe flow of water
number of parameters, some of which difficult to measure. through the xylem using Darcy’s equation. Below ground, flow is
In this study we use a model that couples water flow in soil to described by:
the flow in the xylem of vegetation, both below and above ground.     
The model builds on that presented by Amenu and Kumar [27]. We @hx @ @hx
qg Ss  kp ðhx Þ þ1 ¼ ksr ðzÞ f ðhÞðhs  hx Þ; ð5Þ
use this model to show the key role of root water compensation in @t @z @z
modulating transpiration in a woodland growing on a duplex soil. where hx ½L is the xylem pressure head, Ss ½M1 LT 2  is storage within
Specifically, we show (i) that root water compensation is able to the xylem, and kp ½LT 1  is the spatially averaged axial hydraulic
explain the sustained transpiration rates observed across several conductivity of the xylem.
relatively dry months and (ii) how the trees partition their water Above ground, flow through the xylem is driven by:
demand from various soil layers under scenarios with different     
root depths. @hx @ Ax @hx
qg Ss  kp ðhx Þ þ1 ¼ 0; ð6Þ
@t @z As @z
2. Model description where Ax is the average xylem cross-sectional area of all the plants
present in the ground area As . Plants are assumed to transpire only
We used a one-dimensional model, thereby taking a big-leaf from the top of the canopy so that there are no sinks along the
approximation, according to which spatial variability of the canopy above-ground xylem; transpiration is a function of solar radiation,
is aggregated in the value of parameters estimated at the stand air temperature, vapor pressure deficit, and leaf water potential,
level. The model describes water flow in the soil and xylem, both as explained in Appendix B. A detailed derivation of Eq. (6) is
below and above ground, with the exchange of water between soil presented in Appendix C.
and roots dependent on the water potential difference between the Xylem conductivity, kp , is a function of xylem potential, accord-
two. ing to [38]:
Water flow in variably saturated soils is described using the  
Richards equation with a sink term for root water uptake: 1
kp ¼ kpmax 1  : ð7Þ

  1 þ expðap ðqghx  bp ÞÞ
@h @hs @ @hs
 kðhs Þ þ1 ¼ ksr ðzÞ f ðhÞðhx  hs Þ; ð1Þ Although in general kp can be different for below- and above-
@hs @t @z @z
ground xylem [34], we assumed here a single value for kp in Eqs.
where h ½L3 L3  is volumetric water content, hs ½L is the pressure (5) and (6).
head in the soil, k ½LT 1  is soil hydraulic conductivity, z ½L is the
vertical coordinate (positive upwards), ksr ½L1 T 1  is the soil-root 3. Case study
conductance, f ðhÞ ½ is a water-stress function reducing root water
uptake, and hx ½L is the pressure head in the root xylem. We use the The data reported here have been presented and analyzed in
relationship given by van Genuchten [36] to relate hs to k and h previous studies [12,39–43]. We refer the reader to these studies
(Appendix A). for more complete details on the site. Only a brief description of
Since the exchange of water between soil and roots in different the site and data used in our analysis is given here.
soil layers depends on the amount of roots present in those layers,
the parameter ksr is modeled as a function of root density distribu- 3.1. Site description
tion as:
The site is located at latitude 33 390 4100 S and longitude 150
rðzÞ 0
46 5700 E in New South Wales, Australia. The nearest weather
ksr ðzÞ ¼ R 0 ksrt ; ð2Þ
d
rðzÞdz station with rainfall and temperature data is located at the Royal
Australian Air Force base in Richmond (Australian Bureau of Mete-
where ksrt ½T 1  is total soil-to-root radial conductance, rðzÞ is the orology, station 067105). The long term statistics (1993–2013)
root mass distribution as a function of depth, and d is root depth. show that the average daily minimum and maximum tempera-
The function rðzÞ is assumed as [16]: tures are 10  C and 24  C, with January being the hottest month
P. Verma et al. / Advances in Water Resources 74 (2014) 91–101 93

(average minimum and maximum temperatures are 18  C and distribution was estimated from trenches, with root masses deter-
30  C) and July the coldest (average daily minimum and maximum mined up to a depth of 1:5 m. The root biomass in the first 0:4 m
temperatures are 4  C and 17  C). Mean annual rainfall is about varied between 40% and 80% of the total root biomass, according
730 mm and the wettest month is February with average rainfall to measurements in four trenches [40]; an estimate of root distribu-
of about 125 mm. tion from these measurements is reported in Appendix D. Sap flux
Rainfall, solar radiation, air temperature and humidity were data were collected using the heat ratio technique at half-hour
measured at the field site using a weather station; data from Janu- intervals (Fig. 1); sensors were installed at about 1:3 m from the
ary 1st to June 4th in 2007 were available for this study (Fig. 1). ground. Details on the sap flux measurements, including number
Total rainfall in this period was about 500 mm. of monitored trees, number of sensors per tree, corrections associ-
The soil consists of two layers: the first is dominated by sand ated with wounding, estimation of zero flow, and scaling from trees
(sand to loamy sand) up to a depth of 0:8 m, and the second is clay to stand, are reported in Zeppel et al. [12,39]. Gaps in the sap-flux
with a low amount of sand. The hydraulic properties for these soil measurements totaled 13 days. Discrete measurements of leaf
types are reported in [40] and values of the parameters used in the water potential and soil moisture profiles to a depth of 5 m, span-
model are listed in Table 1. ning the two-year period 2007–2009, are reported in Zeppel et al.
The vegetation is primarily dominated by Eucalyptus parramatt- [12] and Yunusa et al. [42] respectively. These soil moisture
ensis C.A. Hall (Parramatta Red Gum) and Angophora bakeri E.C. Hall measurements were used to parameterize the model, whereas the
(narrow-leaved apple). The trees are about 14 m tall. Leaf area leaf water potential and fine scale measurements of sap flux were
index (LAI) was estimated between 1:3 and 1:9. Root biomass used to evaluate the model performance.

Fig. 1. Observed air temperature (T a ) and rainfall (R), solar radiation (S), vapor pressure deficit (VPD), and sap flux (E) between January 1st and June 7th, 2007.

Table 1
List of soil parameters.

Parameters Units Sand Clay Description Reference


Measured
ks m s1 3:45  105 1:94  107 Saturated hydraulic conductivity [40]

From literature
hs – 0:47 0:55 Saturated volumetric soil moisture content [44]
hr – 0:045 0:068 Residual volumetric soil moisture content [44]
a m1 14:5 0:8 Soil hydraulic parameter [44]
n – 2:4 1:5 Soil hydraulic parameter [44]
l – 0:5 0:5 Soil hydraulic parameter [44]
h1 – 0:05 0:08 Wilting point [44]
Estimated
h2 – 0:09 0:12 Root water uptake reduction parameter
94 P. Verma et al. / Advances in Water Resources 74 (2014) 91–101

Transpiration on available data [12,39–43], we used a soil profile consisting of


a sand layer of 0:8 m above a clay layer of 4:2 m. The water reten-
tion and the hydraulic properties of the two soil types were taken
from Macinnis-Ng et al. [40] and Carsel and Parrish [44] (Table 1).
We set the pressure head at the bottom boundary to be constant
and equal to 6:09 m, corresponding to a volumetric water content
of 0:28. This is consistent with discrete soil moisture observations
reported by Yunusa et al. [42], which showed that the water con-
tent in the clay layer at a depth of 5 m exhibited small fluctuations
14 m around this value. The measured rainfall rate was used as a flux
boundary condition at the surface. As an initial condition, we
XYLEM assumed the pressure head in the sand layer to be constant and
equal to 0:402 m, corresponding to a volumetric moisture con-
Rainfall tent of 0:08; this is in agreement with estimates from soil moisture
z measurements. In the clay layer, we assumed that the pressure
head below an elevation of 3 m was constant and equal to
6:09 m; this is in agreement with the soil moisture measure-
sand 0.8 m ments reported by Yunusa et al. [42]. Between elevations 3 and
4:2 m, the pressure head was interpolated linearly from 6:09 to
water exchange 0:402 m.
d
between root The plant domain consists of above and below-ground xylem. In
soil and roots distribution the soil–plant–atmosphere model presented by Zeppel et al. [12]
clay 4.2 m and applied to the same site, the Authors used a maximum root
depth of 3:2 m, stressing that, although measurements showed
that most of the roots were in the top 1:5 m, the model was not
SOIL able to simulate the sustained transpiration rates unless a depth
of at least 3 m was assumed. In accordance with Zeppel et al.
[12], we also used a root depth of 3:2 m. The boundary conditions
for the plant domain are a zero-flux condition at the bottom of the
Fig. 2. Schematic diagram of the setting used in the simulations. roots and, above ground, transpiration rates define the flux at the
top of the canopy. We used the Penman–Monteith equation cou-
3.2. Numerical simulations pled to the Jarvis formulation for canopy stomatal conductance
to simulate transpiration. Since positive sap flux was often mea-
COMSOL Multiphysics (Ver. 4.1; http://www.comsol.com/) was sured during the night, and according to the Jarvis formulation sto-
used to solve the system of partial differential equations. matal conductance is zero in the absence of solar radiation, we
Fig. 2 shows a schematic of the soil and the plant domains as included nocturnal transpiration using an empirical function based
well as the physical dimensions used in the simulations. Based on the measurements reported by Zeppel et al. [39] (see details in

Table 2
List of vegetation parameters.

Parameter Units Value Description Reference


Measured
Ax =As – 8:62  104 Xylem cross-sectional area and site surface ratio [43]
LAI – 1:5 Leaf area index [12]
qz – 9 Root distribution parameter [40]
From literature
ap Pa1 2  106 Xylem cavitation parameter [38,45]
bp Pa 1:5  106 Xylem cavitation parameter [38,45]
Cp J m3 K1 1200 Heat capacity of air [27]
gb m s1 2  102 Leaf boundary layer conductance [46]
ga ms 1
2  102 Aerodynamic conductance [46]
kpmax m s1 1  105 Xylem conductivity [38,47]
kr m2 W1 5  103 Jarvis radiation parameter [46]
kt K2 1:6  103 Jarvis temperature parameter [46]
k J m3 2:51  109 Latent heat of vaporization [27]
c Pa K1 66:7 Psychrometric constant [27]

Estimated
g smax m s1 10  103 Maximum leaf stomatal conductance
hx50 m 130 Jarvis leaf water potential parameter
kd Pa1 1:1  103 Jarvis vapor pressure deficit parameter
ksrt s1 7:2  1010 Total soil-to-root radial conductance
nl – 2 Jarvis leaf water potential parameter
Ss Pa1 1:1  1011 Xylem storage
T opt K 289:15 Jarvis temperature parameter
P. Verma et al. / Advances in Water Resources 74 (2014) 91–101 95

0.16 0.16

0.12 0.12
E (mm h )
−1

0.08 0.08

0.04 0.04

14 16 18 20 101 103 105 107

Fig. 3. Top: Comparison between measured (circles) and modeled (lines) sap flux rates at 1:3 m above ground for two different time periods. Bottom: corresponding root
water uptake rates, S [103 d1], profiles in the top 0:8 m from the surface. The total root depth is 3:2 m.

Appendix B). As an initial condition, we assume that the xylem 10 mm s1 (about 410 mmol m2 s1), consistent with the maxi-
pressure head decreased linearly from 6:09 m at the bottom of mum stomatal conductance observed at the site [48]. The parame-
the roots to 23:3 m at the top of the canopy. ters ksrt ; Ss , and hx50 were randomly varied, within an interval of
The list of xylem parameters are presented in Table 2. The reasonable values, until a combination of parameters producing a
values of most of the parameters were taken from either measure- satisfactory fit between observations and model predictions was
ments at the site or from the literature. The parameters h2 and nl achieved.
were estimated within the range of values found in the literature;
T opt was estimated to be about the annual average daily tempera- 3.3. Results
ture at the site, i.e., 16  C. The parameter kd was estimated at
1:1  103 Pa1, similar to the value used in Daly et al. [46] Measured transpiration rates shown in Fig. 1 present daily
(i.e., 0:8  103 Pa1). The value of g max was assumed to be cycles with very similar magnitudes across several months,

E
E =0.18+0.90 E mod
mod meas 160 E
2 meas
R =0.70
1.5
Transpiration depth (mm)

120
(mm d−1)

1
80
mod
E

0.5
40

0.5 1 1.5 40 80 120


−1 DOY
E (mm d )
meas

Fig. 4. Left: Comparison between measured (Emeas ) and modeled (Emod ) daily sap flux rates at 1:3 m above ground, excluding fluxes during nights. Right: cumulative plot of
measured (dashed line) and modeled (continuous line) sap fluxes.
96 P. Verma et al. / Advances in Water Resources 74 (2014) 91–101

180

Transpiration depth (mm)


−1 120
h (MPa)
x

3.2 m
60 2.5 m
−2 2m
1.5 m
0.9 m

14 16 18 20 0
DOY
0

Leaf water potential (MPa)


−2

−1
h (MPa)

−4
x

−2 −6

50 100 150
DOY
101 103 105 107
DOY Fig. 6. Modeled cumulative transpiration rates (top) and daily minimum leaf water
DOY 14 potentials (bottom) obtained with root depths decreasing from 3:2 to 0:9 m.
DOY 18
12 DOY 101
DOY 105 between daily patterns of measured and modeled sap flux at
1:3 m above ground. Two periods of the year are shown: a dry per-
iod in summer, between days 13 and 20, and a wetter period in
autumn, between days 100 and 107. The patterns of root water
Elevation (m)

8
uptake generating these transpiration rates are also shown in
Fig. 3. In the driest part of the simulated period, between days 13
and 20, when soil near the surface becomes very dry, roots were
able to compensate for this lack of soil water content by increasing
4 water uptake from deeper soil layers. Root water compensation
was also present, although not so clearly evident, during wetter
periods. In these periods, the majority of water for transpiration
was provided by soil near the surface, where most of the roots
are, but part of the transpired water was always supported by
−2 −1 0
h (MPa) deeper soil layers.
x
Rates of daily sap flux, excluding nights, were also satisfactorily
Fig. 5. Modeled leaf water potential (continuous line) and root water potential at reproduced by the model (Fig. 4), with NSE equal to 0:58. The coef-
the surface (dashed line) in two different periods. The panel at the bottom shows ficient of determination between modeled and observed data was
the xylem potential as a function of height above ground at 12 pm in four different
0:70, with the slope of the least-squares regression line being 0:90
days.
(with 95% confidence interval ranging from 0:80 to 1:00) with an
intercept of 0:17 (with 95% confidence interval between 0:06
not-with-standing the reductions in solar radiation, temperature and 0:29). Fig. 4 also shows the cumulative amount of water tran-
and vapor pressure deficit as well as variability of rainfall across spired by the vegetation and that obtained from the model during
the study period. The model was able to capture these sustained the simulated period.
transpiration rates. Examples of modeled water potentials in the xylem are shown
The Nash–Sutcliffe Efficiency (NSE) coefficient for the whole in Fig. 5. The modeled leaf water potential oscillated around
series of data was 0:86, indicating an acceptable agreement 0:2 MPa at pre-dawn and reached values of about 3 MPa in
between modeled and measured sap flux. Since modeling transpi- the middle of the day. These values are consistent with the leaf
ration during the night was not the focus of this work, we removed water potentials measured by Zeppel et al. [12] on four days during
the data corresponding to nights; the NSE in this case remained the period September–December; measured pre-dawn values for
high, but was reduced to 0:71. Fig. 3 shows the comparison E. parramattensis and A. bakeri were between 0:18 and
P. Verma et al. / Advances in Water Resources 74 (2014) 91–101 97

1:77 MPa, while minimum leaf water potentials were between modeled the root water uptake when root depths lower than
1:81 and 3:60 MPa. In the same study, Zeppel et al. [12] used 3:2 m are assumed. We recall that in the model used by Zeppel
in their soil–plant–atmosphere model a value of 2:8 MPa as a et al. [12], which did not account for root water compensation, a
minimum sustainable leaf water potential, which is consistent minimum root depth of 3 m was required to satisfactorily describe
with the values obtained in our model. The modelled difference the observed sap flow rates. Eucalyptus species are generally ever-
between root and leaf water potentials during the night were green and require deep root systems; other species (e.g., drought
due to both differences in elevations (14 m ¼ 0:14 MPa between deciduous species) might require shallower root depths. We now
soil surface and leaves) and transpiration during the night. The consider root depths decreasing from a maximum of d ¼ 3:2 m to
gradient in water potentials between roots and leaves were suffi- a minimum of d ¼ 0:9 m, to establish whether different root depths
cient to generate a flux of water from the soil to the atmosphere, in our model are able to maintain the observed sap fluxes. We thus
such that the model did not show fluxes of water from roots to assumed that the root distribution had the shape as in Eq. (3),
the soil possibly associated with hydraulic redistribution. Fig. 5 modifying only the value of d; since we used Eq. (2), this assump-
also shows variation of above-ground xylem water potential for tion implies that the root mass remained the same in all simula-
different days at 12 pm. The decrease of the xylem water potential tions, but it was distributed differently along the root depth.
with height is non-linear because of the reduction in xylem According to the model results, the water transpired over the
conductivity (Eq. 7) caused by low leaf water potentials. period under analysis decreased with shallower root depths
(Fig. 6). Similar transpiration rates were obtained with root depths
from 3:2 to 2 m, while transpiration declined more noticeably with
4. Root water uptake with different root depths depths lower than about 2 m. Specifically, when assuming that the
roots were almost entirely in the sandy layer (i.e., d ¼ 0:9 m), the
Previous studies and the sap flux data in Fig. 1 show that trees total transpiration after 155 days was about 24 mm lower than in
at the site are able to maintain sustained daily sap-flux rates across the case with d ¼ 3:2 m. In order to maintain such high transpira-
several months; the reason for this has been attributed to the con- tion rates even with shallower root depths, the vegetation needs to
tinuous presence of water near the interface between the sand and lower the xylem water potential. As shown in Fig. 6, the minimum
clay layers, which acts as a buffer during dry periods [12]. Thus, leaf water potential that the trees needed to sustain became much
when the sand becomes dry near the surface, the trees are able lower as root depth decreased. The values of leaf water potential
to increase their water uptake from deeper soil layers, even though obtained with root depths lower than about 2 m were very low
a lower percentage of roots are present at these depths. Since data and unrealistic when compared to measurements at the site [12].
on root density were available only to a depth of 1:5 m, we For all root depths, the daily minimum leaf water potential was

0.16 0.16

0.12 0.12
E (mm h )
−1

0.08 0.08

0.04 0.04

14 16 18 20 101 103 105 107

Fig. 7. Top: Modeled sap flux rates at 1:3 m from the ground for two different time periods using a total root depth of 0:9 m. Bottom: corresponding root water uptake rates,
S [103 d1], profiles in the top 0:8 m from the surface.
98 P. Verma et al. / Advances in Water Resources 74 (2014) 91–101

below hx50 for 91% of the days. With d ¼ 3:2 m, the daily averaged comparison to the case with deeper root depths. Even in wetter
leaf water potential stayed above hx50 during the entire period periods, the shallow rooted vegetation is more dependent on
simulated; as d decreased, the percentage of days in which the rainfall. This is for example shown by the rapid water uptake near
daily averaged leaf water potential went below hx50 increased from the surface occurring towards the end of the day 104, when 1 mm
1.3%, for d ¼ 2:5 m, to 11%, when d was 0:9 m. of water fell in 30 min (Fig. 7). This small event did not cause any
The reduction in transpiration with shallower root depths is significant change in the root water uptake when the root depth
evident by comparing the results in Fig. 7 with those in Fig. 3. With was set to 3:2 m (Fig. 3).
shallower root depths, the vegetation was more reliant on soil Most of the root water uptake occurred within the first 0:8 m of
moisture near the surface. Larger water uptake rates near the sur- soil. Fig. 8 shows how daily root water uptake is partitioned among
face after rainfall events caused a quick drying of the sandy layer. various parts of the root system comparing different root depths.
With water uptake occurring mostly from the sand layer and the With d ¼ 3:2 m, about 15% of the daily root water uptake consis-
high hydraulic conductivity of the sand, soil moisture content tently came from soil layers deeper than 0:6 m, where about 15%
declined near the surface inducing water stress in the vegetation. of the roots are; during dry periods, the percentage of water from
Root water compensation was still evident during dry periods, these depths increased, reaching about 30% in the driest part of the
but the lower moisture available reduced the water uptake in period analyzed. When reducing the root depth, the roots between
0:3 and 0:6 m became more important during dry periods. They
contributed at times more than 50% of the total daily rott water
d = 3.2 m uptake, even though less than 20% of the roots were at these
90 depths.
0-0.1 m (27% of roots)
To see the effect of lower rainfall rates on the partitioning of
root water uptake, we ran the same simulations using as input
70% of the rainfall rates measured at the site (i.e., a total rainfall
of 350 mm). The total transpiration rates in this case were slightly
60
lower than those obtained before, and these transpiration rates
% RWU

0.1-0.3m (34%)
were achieved thanks to lower leaf water potentials (not shown).
The partition of root water uptake from different depths did not
differ considerable from what shown in Fig. 8.
30
0.3-0.6 m (24%)

0.6-0.9 m (9%) 5. Conclusions


<0.9 m (6%)

d=2m
Root water compensation is the mechanism through which veg-
90 0-0.1 m (39% of roots) etation adjusts the depth of root water uptake based on soil water
availability, thereby favoring uptake from areas of the soil that
have higher water potentials even when the root density is rela-
tively low in these areas. We present here a modeling study that
helps show the key role of root water compensation in maintaining
60
sustained transpiration rates in an Australian woodland growing
% RWU

on a duplex soil.
0.1-0.3 m (39%)
Using a one dimensional model that couples water flow in the
soil to that in the xylem, we were able to reproduce field observa-
30 tions of transpiration rates over a period of several months. The
root depth in this simulation was initially assumed to be 3:2 m;
0.6-0.9 m (4%) this value was adopted in a previous study at the same site [12].
0.3-0.6 m (17%) <0.9 m (1%) Root water uptake from the model suggested that root water com-
pensation was a key mechanism that allowed trees to maintain
stable transpiration rates during dry periods of the year; root water
d = 1.5 m
90 0-0.1 m (49% of roots)
compensation was also present during relatively wet periods. The
roots below a depth of 0.6 m, which total about 15% of the whole
root biomass, typically provided 15% of the daily transpiration, but
reached 30% during the driest part of the analyzed period.
Scenarios with different root depths, from 3:2 to 0:9 m, showed
60
that in all cases root water compensation was actively involved in
% RWU

sustaining transpiration rates, which anyway reduced considerably


for root depths shallower than about 2 m. For such depths, transpi-
ration rates were sustained by decreasing leaf water potentials,
0.1-0.3 m (38%)
30 which were much lower than what was experimentally observed.
The model suggested that the vegetation with shallower root sys-
tem is more responsive to rainfall events; in this case the soil sur-
0.3-0.6 m (11%) 0.6-0.9 m (~1%) face becomes drier and vegetation is more prone to experience
<0.9 m (~1%)
water stress, thereby reducing transpiration rates. Root depths
50 100 150 shallower than 2 m, with a percentage of root biomass below
DOY
0:6 m of less than about 5%, reduced transpiration rates because
Fig. 8. Percentage of daily total root water uptake (RWU) from various parts of the of water stress and led to leaf water potentials much lower than
root system for three different root depths, d. field observations. In those scenarios, roots deeper than 0:3 m
P. Verma et al. / Advances in Water Resources 74 (2014) 91–101 99

 
provided up to 70% of the daily transpiration even though root bio- 17:27ðT a  273:15Þ
esat ¼ 611 exp ; ðB:3Þ
mass at these depths was less than 15%. ðT a  35:85Þ
Our results can likely be extended to other areas with duplex
with esat in Pascal.
soils and shallow water table conditions, where large amounts of
The canopy conductance, g c , was modeled as:
water can be reached by the roots. In such conditions, roots can
 
switch their preferential water uptake from near the surface gs gb
gc ¼ LAI; ðB:4Þ
immediately after rainfall events to deeper soil layers, where the gs þ gb
wet soil allows for larger differences between soil and root water
potentials. Our study thus shows that duplex soils might provide where LAI ½ is the leaf area index, g b ½LT 1  is the leaf boundary
energetically favorable conditions for root water uptake near the layer conductance per unit leaf area and g s ½LT 1  is the stomatal
interface between the layers, thereby allowing vegetation to main- conductance. Assuming that CO2 concentrations remain constant
tain sustained transpiration rates not-with-standing intermittent and do not affect the stomatal conductance, g s can be modeled as
wet and dry periods. [51]:
g s ¼ g smax f ðSÞf ðT a Þf ðDÞf ðhxleaf Þ; ðB:5Þ
Acknowledgements 1
where g smax ½LT  is the maximum stomatal conductance per unit
The authors thank Melanie J. B. Zeppel for providing the data leaf area, and f ðSÞ ½; f ðT a Þ ½; f ðDÞ ½, and f ðhxleaf Þ ½ are empiri-
used in the study. P.V., D.E. and E.D. acknowledge the support of cal functions varying between 0 and 1, which scale g smax depending
the Australian Research Council (ARC) and the Australian National on solar radiation, air temperature, vapor pressure deficit and leaf
Water Commission (NWC) through Program 4 (Groundwater Veg- water potential. These functions are commonly written as:
etation Atmosphere Interactions) of the National Centre for f ðSÞ ¼ 1  expðkr SÞ; ðB:6Þ
Groundwater Research and Training. P.V. thanks the Director of
the National Environmental Engineering Research Institute, Nag- f ðT a Þ ¼ 1  kt ðT a  T opt Þ2 ; ðB:7Þ
pur, India, for granting leave to pursue his PhD studies at Monash
University. 1
f ðDÞ ¼ ; ðB:8Þ
ð1 þ D kd Þ
Appendix A. Water retention curves
  n 1
hxleaf l
In unsaturated conditions, the soil moisture (h) and unsaturated f ðhxleaf Þ ¼ 1 þ ; ðB:9Þ
hx50
hydraulic conductivity (k) vary with soil pressure head (hs ). The
relationship between these variables are modeled as [36]: where kr ; kt ; kd and nl are empirical constants, T opt is the air temper-
( hs hr ature at which f ðTÞ is 1, and hx50 is the leaf water potential at which
hr þ ½1þjahs jn 
m if hs < 0;
h¼ ðA:1Þ f ðhxleaf Þ is 0:5.
hs if hs P 0; We model the night time transpiration as:
8 En ¼ Emax f ðT a Þf ðDÞf ðhxleaf Þ; ðB:10Þ
>  l    1=m m 2
<
ks hhh r
1  1  hhh r
if hs < 0; 1
where Emax ½LT  is the maximum night time transpiration assumed
k¼ s hr s hr ðA:2Þ
>
: to be equal to 1  109 m s1 based on the observed night time sap
ks if hs P 0;
flux.
where hr ½L L  is the residual water content, hs ½L3 L3  is the water
3 3

content at saturation, ks ½LT 1  is the soil hydraulic conductivity at Appendix C. Water flow in above ground xylem
saturation and a½L1 ; l½; n½ are parameters dependent on soil
type; the parameter m½ is equal to 1  1=n. Above ground, for a single plant, the water flow through the
xylem can be written as [52,53]:
Appendix B. Penman–Monteith equation     
@hx 1 @ @hx
qg Ss  kp ðhx ; zÞAðzÞ þ1
@t AðzÞ @z @z
The Penman–Monteith equation reads:
1
  ¼ sðzÞ Ed ðz; t; hx Þ; ðC:1Þ
Q n D þ C p Dg a AðzÞ
Ep ¼ g; ðB:1Þ
k½Dg c þ cðg c þ g a Þ c
where AðzÞ ½L2  is the cross-sectional area of the above ground
1 3
where Ep ½LT  is the transpiration rate, Q n ½MT  is the net radia- xylem, kp ½LT 1  is the xylem conductivity per unit of xylem cross-
tion, D ½ML1 T 2 K 1  is the slope of the saturation vapor pressure sectional area, sðzÞ ½L is the leaf area per unit of stem length, and
curve at a given air temperature, C p ½ML1 T 2 K 1  is the heat capac- Ed ðz; t; hx Þ ½LT 1  is the transpiration flux density. For simplicity, we
ity of air at constant pressure, D ½ML1 T 2  is the vapour pressure assume a uniform xylem cross sectional area above ground, and
deficit, k ½ML1 T 2  is the latent heat of vaporization, g a ½LT 1  is thus AðzÞ becomes a constant. Therefore, AðzÞ on the l.h.s. of
the aerodynamic conductance, c ½ML1 T 2 K 1  is the psychrometric Eq. (C.1) at the numerator and denominator cancel each other. Since
constant and g c ½LT 1  is the canopy conductance. The net radiation we do not consider a canopy structure, the term on the r.h.s of
is calculated as the 70% of the total radiation, S [49]. The slope of the Eq. (C.1) becomes zero.
saturation vapor pressure curve, D, is obtained as [50]: Because Eqs. (1) and (5) define transpiration per unit of ground
" # area, we need to re-scale Eq. (C.1) from the xylem cross-sectional
4098 area of a single plant to the unit of ground area. This is obtained
D¼ esat ; ðB:2Þ by re-scaling the axial hydraulic conductivity, kp , which is multi-
ðT a  35:85Þ2
plied by the ratio Ax =As , where Ax is the average xylem cross-
where esat ½ML1 T 2  is the saturation vapor pressure at a given air sectional area of all the plants in the ground area As . Accordingly,
temperature T a ½K, and given by the equation [50]: Eq. (C.1) becomes Eq. (6).
100 P. Verma et al. / Advances in Water Resources 74 (2014) 91–101

0 [11] Orellana F, Verma P, Loheide SPI, Daly E. Monitoring and modeling water-
vegetation interaction in groundwater-dependent ecosystems. Rev Geophys
2012;50(RG3003):1–24. http://dx.doi.org/10.1029/2011RG000383.
[12] Zeppel M, MacInnis-Ng C, Palmer A, Taylor D, Whitley R, Fuentes S, Yunusa I,
Williams M, Eamus D. An analysis of the sensitivity of sap flux to soil and plant
variables assessed for an Australian woodland using a soil–plant–atmosphere
−1
model. Funct Plant Biol 2008;35(6):509–20. http://dx.doi.org/10.1071/FP08114.
[13] Jarvis NJ. A simple empirical model of root water-uptake. J Hydrol
1989;107(1–4):57–72. http://dx.doi.org/10.1016/0022-1694(89)90050-4.
Depth(m)

[14] Javaux M, Couvreur V, Vander Borght J, Vereecken H. Root water uptake: from
three-dimensional biophysical processes to macroscopic modeling approaches.
−2 Vadose Zone J 2013;12(4). http://dx.doi.org/10.2136/vzj2013.02.0042.
[15] Jarvis NJ. Simple physics-based models of compensatory plant water uptake:
concepts and eco-hydrological consequences. Hydrol Earth Syst Sci
2011;15(11):3431–46. http://dx.doi.org/10.5194/hess-15-3431-2011.
[16] Vrugt JA, Van Wijk MT, Hopmans JW, imunek J. One-, two-, and three-
dimensional root water uptake functions for transient modeling. Water Resour
−3 Res 2001;37(10):2457–70. http://dx.doi.org/10.1029/2000WR000027.
[17] Vrugt JA, Hopmans JW, Simunek J. Calibration of a two-dimensional root water
uptake model. Soil Sci Soc Am J 2001;65(4):1027–37. http://dx.doi.org/
1 2 3 10.2136/sssaj2001.6541027x.
root density (m )
−1 [18] Skaggs T, van Genuchten M, Shouse P, Poss J. Macroscopic approaches to root
water uptake as a function of water and salinity stress. Agric Water Manage
Fig. D.9. Modeled root distribution with d ¼ 3:2 m (continuous line) and measured 2006;86(1–2):140–9. http://dx.doi.org/10.1016/j.agwat.2006.06.005.
distribution (circles) [40]. [19] Li KY, De Jong R, Boisvert JB. An exponential root-water-uptake model with
water stress compensation. J Hydrol 2001;252(1–4):189–204. http://
dx.doi.org/10.1016/S0022-1694(01)00456-5.
Appendix D. Estimated root distribution [20] Yadav BK, Mathur S, Siebel MA. Soil moisture dynamics modeling considering
the root compensation mechanism for water uptake by plants. J Hydrol Eng
2009;14(9):913–22. http://dx.doi.org/10.1061/(ASCE)HE.1943-5584.0000066.
The root distribution when the root depth equals 3:2 m was [21] Simunek J, Hopmans JW. Modeling compensated root water and nutrient
estimated from the data presented in Fig. 4 of Macinnis-Ng et al. uptake. Ecol Modell 2009;220(4):505–21. http://dx.doi.org/10.1016/
j.ecolmodel.2008.11.004.
[40]. The percentage of roots at 0:1; 0:3; 0:5; 1:0, and 1:5 m was [22] Siqueira M, Katul G, Porporato A. Onset of water stress, hysteresis in plant
measured in four trenches. We averaged the results from the four conductance, and hydraulic lift: Scaling soil water dynamics from millimeters
trenches and assumed that the percentage of roots at 0:1 m was to meters. Water Resour Res 2008;44(W01432):1–14. http://dx.doi.org/
10.1029/2007wr006094.
distributed over the first 0:15 m of soil; the value at 0:3 m referred [23] Siqueira M, Katul G, Porporato A. Soil moisture feedbacks on convection
to the soil between 0:15 and 0:40 m, the value at 0:5 m referred to triggers: the role of soil-plant hydrodynamics. J Hydrometeorol
the soil between 0:40 and 0:75 m, the value at 1:0 m referred to the 2009;10(1):96–112. http://dx.doi.org/10.1175/2008jhm1027.1.
[24] Tron S, Laio F, Ridolfi L. Plant water uptake strategies to cope with stochastic
soil between 0:75 and 1:25 m, and the value at 1:5 m referred to
rainfall. Adv Water Resour 2013;53:118–30. http://dx.doi.org/10.1016/
the depths between 1:25 and 1:75 m. Dividing the values of the j.advwatres.2012.10.007.
averaged percentage of roots by the respective depth interval gives [25] Vogel T, Dohnal M, Dusek J, Votrubova J, Tesar M. Macroscopic modeling of
plant water uptake in a forest stand involving root-mediated soil water
and estimation of the root density.
redistribution. Vadose Zone J 2013;12(1). http://dx.doi.org/10.2136/
The parameter qz of the root distribution (Eq. 2) with d ¼ 3:2 m vzj2012.0154.
was chosen to fit the root estimated from these calculation, as [26] Mendel M, Hergarten S, Neugebauer HJ. On a better understanding of
shown in Fig. D.9. hydraulic lift: a numerical study. Water Resour Res 2002;38(10):11–110.
http://dx.doi.org/10.1029/2001WR000911.
[27] Amenu GG, Kumar P. A model for hydraulic redistribution incorporating
References coupled soil-root moisture transport. Hydrol Earth Syst Sci 2008;12(1):55–74.
http://dx.doi.org/10.5194/hess-12-55-2008.
[28] Quijano JC, Kumar P, Drewry DT, Goldstein A, Misson L. Competitive and
[1] Bonan G. Ecological climatology: concepts and applications. Cambridge,
mutualistic dependencies in multispecies vegetation dynamics enabled by
UK: Cambridge University Press; 2002.
hydraulic redistribution. Water Resour Res 2012;48(W05518):1–22. http://
[2] Arya LM, Blake GR, Farrell DA. Field study of soil-water depletion patterns in
dx.doi.org/10.1029/2011wr011416.
presence of growing soybean roots. 3. Rooting characteristics and root
[29] Gou S, Miller G. A groundwater–soil–plant–atmosphere continuum approach
extraction of soil-water. Soil Sci Soc Am J 1975;39(3):437–44. http://
for modelling water stress, uptake, and hydraulic redistribution in phreatophytic
dx.doi.org/10.2136/sssaj1975.03615995003900030023x.
vegetation. Ecohydrology 2013;7(3):1029–41. http://dx.doi.org/10.1002/
[3] Nnyamah JU, Black TA. Rates and patterns of water-uptake in a Douglas-fir
eco.1427.
forest. Soil Sci Soc Am J 1977;41(5):972–9. http://dx.doi.org/10.2136/
[30] van Lier QD, van Dam JC, Metselaar K, de Jong R, Duijnisveld WHM.
sssaj1977.03615995004100050033x.
Macroscopic root water uptake distribution using a matric flux potential
[4] Green SR, Clothier BE. Root water uptake by kiwifruit vines following partial
approach. Vadose Zone J 2008;7(3):1065–78. http://dx.doi.org/10.2136/
wetting of the root zone. Plant Soil 1995;173(2):317–28. http://dx.doi.org/
vzj2007.0083.
10.1007/BF00011470.
[31] Jarvis NJ. Comment on macroscopic root water uptake distribution using a
[5] McLean EH, Ludwig M, Grierson PF. Root hydraulic conductance and aquaporin
matric flux potential approach. Vadose Zone J 2010;9(2):499–502. http://
abundance respond rapidly to partial root-zone drying events in a riparian
dx.doi.org/10.2136/vzj2009.0148.
Melaleuca species. New Phytol 2011;192(3):664–75. http://dx.doi.org/
[32] Doussan C, Pierret A, Garrigues E, Pages L. Water uptake by plant roots: II –
10.1111/j.1469-8137.2011.03834.x.
modelling of water transfer in the soil root-system with explicit account of
[6] Caldwell MM, Dawson TE, Richards JH. Hydraulic lift: consequences of water
flow within the root system – comparison with experiments. Plant Soil
efflux from the roots of plants. Oecologia 1998;113(2):151–61. http://
2006;283(1–2):99–117. http://dx.doi.org/10.1007/s11104-004-7904-z.
dx.doi.org/10.1007/s004420050363.
[33] Bohrer G, Mourad H, Laursen TA, Drewry D, Avissar R, Poggi D, Oren R, Katul
[7] Ryel RJ, Caldwell MM, Yoder CK, Or D, Leffler AJ. Hydraulic redistribution in a
GG. Finite element tree crown hydrodynamics model (FETCH) using porous
stand of Artemisia tridentata: evaluation of benefits to transpiration assessed
media flow within branching elements: a new representation of tree
with a simulation model. Oecologia 2002;130(2):173–84. http://dx.doi.org/
hydrodynamics. Water Resour Res 2005;41(W11404):1–17. http://
10.1007/s004420100794.
dx.doi.org/10.1029/2005wr004181.
[8] Prieto I, Kikvidze Z, Pugnaire FI. Hydraulic lift: Soil processes and transpiration
[34] Janott M, Gayler S, Gessler A, Javaux M, Klier C, Priesack E. A one-dimensional
in the Mediterranean leguminous shrub Retama sphaerocarpa (L.) Boiss. Plant
model of water flow in soil-plant systems based on plant architecture. Plant
Soil 2010;329(1):447–56. http://dx.doi.org/10.1007/s11104-009-0170-3.
Soil 2011;341(1–2):233–56. http://dx.doi.org/10.1007/s11104-010-0639-0.
[9] Li LH, Wang YP, Yu Q, Pak B, Eamus D, Yan JH, van Gorsel E, Baker IT. Improving
[35] Bittner S, Legner N, Beese F, Priesack E. Individual tree branch-level simulation
the responses of the Australian community land surface model (CABLE) to
of light attenuation and water flow of three F. sylvatica L. trees. J Geophys Res-
seasonal drought. J Geophys Res-Biogeosci 2012;117. http://dx.doi.org/
Biogeosci 2012;117(G01037):1–17. http://dx.doi.org/10.1029/2011JG001780.
10.1029/2012jg002038.
[36] van Genuchten MT. Closed-form equation for predicting the hydraulic
[10] Guswa AJ. Canopy vs. roots: production and destruction of variability in soil
conductivity of unsaturated soils. Soil Sci Soc Am J 1980;44(5):892–8. http://
moisture and hydrologic fluxes. Vadose Zone J 2012;11(3). http://dx.doi.org/
dx.doi.org/10.2136/sssaj1980.03615995004400050002x.
10.2136/vzj2011.0159.
P. Verma et al. / Advances in Water Resources 74 (2014) 91–101 101

[37] Feddes RA, Kowalik P, Kolinska-Malinka K, Zaradny H. Simulation of field [46] Daly E, Porporato A, Rodriguez-Iturbe I. Coupled dynamics of photosynthesis,
water uptake by plants using a soil water dependent root extraction function. J transpiration, and soil water balance. Part I: upscaling from hourly to daily
Hydrol 1976;31(1–2):13–26. http://dx.doi.org/10.1016/0022-1694(76)90017-2. level. J Hydrometeorol 2004;5(3):546–58. http://dx.doi.org/10.1175/1525-
[38] Pammenter NW, Vander Willigen C. A mathematical and statistical analysis of 7541(2004)005<0546:CDOPTA>2.0.CO;2.
the curves illustrating vulnerability of xylem to cavitation. Tree Physiol [47] Prior LD, Eamus D. Seasonal changes in hydraulic conductance, xylem
1998;18(8–9):589–93. http://dx.doi.org/10.1093/treephys/18.8-9.589. embolism and leaf area in Eucalyptus tetrodonta and Eucalyptus miniata
[39] Zeppel M, Tissue D, Taylor D, Macinnis-Ng C, Eamus D. Rates of nocturnal saplings in a north Australian savanna. Plant Cell Environ 2000;23(9):955–65.
transpiration in two evergreen temperate woodland species with differing http://dx.doi.org/10.1046/j.1365-3040.2000.00612.x.
water-use strategies. Tree Physiol 2010;30(8):988–1000. http://dx.doi.org/ [48] Medlyn BE, Duursma RA, Eamus D, Ellsworth DS, Prentice IC, Barton CVM,
10.1093/treephys/tpq053. Crous KY, De Angelis P, Freeman M, Wingate L. Reconciling the optimal and
[40] Macinnis-Ng CMO, Fuentes S, O’Grady AP, Palmer AR, Taylor D, Whitley RJ, empirical approaches to modelling stomatal conductance. Global Change Biol
Yunusa I, Zeppel MJB, Eamus D. Root biomass distribution and soil properties 2011;17(6):2134–44. http://dx.doi.org/10.1111/j.1365-2486.2010.02375.x.
of an open woodland on a duplex soil. Plant Soil 2010;327(1–2):377–88. [49] Lhomme JP, Elguero E, Chehbouni A, Boulet G. Stomatal control of
http://dx.doi.org/10.1007/s11104-009-0061-7. transpiration: examination of Monteith’s formulation of canopy resistance.
[41] Macinnis-Ng C, Zeppel M, Williams M, Eamus D. Applying a SPA model to Water Resour Res 1998;34(9):2301–8. http://dx.doi.org/10.1029/98WR01339.
examine the impact of climate change on GPP of open woodlands and the [50] Allen RG, Pereira LS, Raes D, Smith M. Crop evapotranspiration - Guidelines for
potential for woody thickening. Ecohydrology 2011;4(3):379–93. http:// computing crop water requirements - FAO irrigation and drainage paper
dx.doi.org/10.1002/eco.138. 56. Rome: FAO; 1998.
[42] Yunusa IA, Zeppel MJ, Fuentes S, Macinnis-Ng CM, Palmer AR, Eamus D. An [51] Jarvis PG. The interpretation of the variations in leaf water potential and
assessment of the water budget for contrasting vegetation covers associated stomatal conductance found in canopies in the field. Philos Trans R Soc London
with waste management. Hydrol Processes 2010;24(9):1149–58. http:// Ser B, Biol Sci 1976;273(927):593–610. http://dx.doi.org/10.1098/rstb.1976.0035.
dx.doi.org/10.1002/hyp.7570. [52] Kumagai T. Modeling water transportation and storage in sapwood – model
[43] Yunusa IAM, Zolfaghar S, Zeppel MJB, Li Z, Palmer AR, Eamus D. Fine root development and validation. Agric For Meteorol 2001;109(2):105–15. http://
biomass and its relationship to evapotranspiration in woody and grassy dx.doi.org/10.1016/s0168-1923(01)00261-1.
vegetation covers for ecological restoration of waste storage and mining [53] Chuang YL, Oren R, Bertozzi AL, Phillips N, Katul GG. The porous media model
landscapes. Ecosystems 2012;15(1):113–27. http://dx.doi.org/10.1007/ for the hydraulic system of a conifer tree: Linking sap flux data to transpiration
s10021-011-9496-9. rate. Ecol Modell 2006;191(3–4):447–68. http://dx.doi.org/10.1016/
[44] Carsel RF, Parrish RS. Developing joint probability-distributions of soil-water j.ecolmodel.2005.03.027.
retention characteristics. Water Resour Res 1988;24(5):755–69. http://
dx.doi.org/10.1029/WR024i005p00755.
[45] Froend RH, Drake PL. Defining phreatophyte response to reduced water
availability: preliminary investigations on the use of xylem cavitation
vulnerability in Banksia woodland species. Aust J Bot 2006;54(2):173–9.
http://dx.doi.org/10.1071/BT05081.

You might also like