You are on page 1of 25

17

Plant Canopies

forests illustrate the environmental controls


17.1  Chapter Summary of canopy fluxes and the partitioning of fluxes
between the canopy and forest floor.
The principles that determine the tempera-
ture, energy balance, and photosynthetic rate
of a leaf also determine those of plant canopies 17.2  Leaf Area Index
when integrated over all leaves in the canopy.
These processes are related to the amount of Leaf area index measures the amount of foliage
leaf area, quantified by leaf area index. The ver- in a plant canopy. Leaf area index is the projected
tical profile of leaf area in the canopy affects the area of leaves per unit area of ground. A square
distribution of radiation in the canopy and the centimeter of ground covered by a leaf with an
absorption of radiation by leaves. With low leaf area of 1 cm2 has a leaf area index of one. The
area index, plants absorb little solar radiation, same area covered by two leaves, one atop the
and the overall surface albedo is largely that other and each with an area of 1 cm2, has a leaf
of soil. The absorption of radiation increases area index of two. Projected, or one-sided, leaf
with greater leaf area index, and surface albedo area is different from total leaf area. For thin,
responds more to the optical properties of flat leaves such as broadleaf trees, total leaf area
foliage rather than soil. The integration of leaf is twice the projected leaf area (both sides of the
processes over the light profile is central to the leaf are included). For needleleaf trees, total leaf
scaling of processes from a leaf to the canopy. area is more than twice the projected leaf area.
The carbon uptake by a canopy is the integra- A leaf area index of two means there are 2 m2
tion of the photosynthetic rates of individual of leaf area (one-sided) covering 1 m2 of ground.
leaves, accounting for variations in light, micro- A typical leaf area index for a productive forest
climate, and foliage nitrogen with depth in the is 4–6 m2 m–2.
canopy. Similarly, canopy conductance is an Cumulative leaf area index increases pro-
aggregate measure of the conductance of indi- gressively with greater depth from the top of
vidual leaves. The profile of leaf area in the can- the canopy. Leaf area index measured below
opy also affects turbulence within the canopy. the canopy is the total leaf area. This leaf area is
The influence of vegetation on surface fluxes typically not distributed uniformly with height.
can be modeled by treating the soil–canopy sys- Many forests have a single layer of foliage in the
tem as an effective bulk surface, or a big leaf. overstory. The oak forest shown in Figure 17.1,
The Penman–Monteith equation applied to can- for example, has an overstory layer at a height
opies is an example of a big-leaf model. Fluxes between 6.5 and 3 m with a cumulative leaf area
of CO2 and evapotranspiration measured over index of 4.6 m2 m–2. The foliage is densest at a

Downloaded from https://www.cambridge.org/core. Danish National Library of Science and Medicine, on 25 Apr 2022 at 11:50:28, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781107339200.018
17.3  Radiative Transfer 265

Solar radiation (%)


(a) (b)
0 20 40 60 80 100
7 7
6 6
5 5

Height (m)
Height (m)
4 4
3 3 Radiation LAI

2 2
1 1
Oak forest
0 0
0.00 0.50 1.00 1.50 0 1 2 3 4 5
Leaf area index Cumulative leaf area index

(c) (d) Solar radiation (%)


0 25 50 75 100
12 12
10 10
8 8

Height (m)
Height (m)

6 6
Radiation LAI
4 4

2 2
Aspen forest
0 0
0.00 0.50 1.00 1.50 2.00 0 1 2 3 4 5 6 7 8
Leaf area index Cumulative leaf area index

Fig. 17.1 Vertical profile of leaves in forests. The top panels show for an oak forest (a) the leaf area profile and (b) the
cumulative leaf area and irradiance as a percentage of that at the top of the canopy. The bottom panels (c–d) show the same
data for an aspen forest with understory. Data from Rauner (1976).

height of about 5 m. Cumulative leaf area index radiative transfer shows that solar radiation
at this height is 3 m2 m–2, about two-thirds of decreases as an exponential function of leaf
the total foliage. In contrast, the aspen forest area index:
shown in Figure 17.1 has a leaf area index of 7.1
m2 m–2. Leaves are found in both an overstory I ↓ (z ) = I ↓0 e − K b L ( z ) (17.1)
situated between 10.5 and 5.5 m height and a
lower understory below 3 m. where I ↓0 is the solar radiation at the top of the
canopy, K b is the light extinction coefficient, L(z )
is the cumulative leaf area index at height z, and
17.3  Radiative Transfer I ↓ (z ) is the irradiance at height z.
The extinction coefficient depends on leaf
Plant canopies have a vertical gradient of sun- orientation and solar zenith angle. These deter-
light related to the amount of leaves. An indi- mine the angle at which the solar beam strikes
vidual leaf absorbs, reflects, or transmits the a leaf. A leaf receives the greatest radiation per
sunlight that strikes it. As the canopy becomes unit surface area when the beam is perpendic-
denser with leaves, more solar radiation is ular to the leaf. The angle at which solar radia-
absorbed or reflected and less is transmitted tion strikes a leaf depends on solar zenith angle.
deeper into the canopy. A  simple model of It also depends on leaf orientation. Some leaves

Downloaded from https://www.cambridge.org/core. Danish National Library of Science and Medicine, on 25 Apr 2022 at 11:50:28, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781107339200.018
266 Plant Canopies

are oriented horizontally while others are verti- index of 7.1 m2 m–2. Leaves are found in both an
cal. Many leaves are oriented randomly so that overstory with a leaf area index of 4.9 m2 m–2
there is an equal probability of orientation in and an understory with a leaf area index of 2.2
any direction (also known as a spherical leaf m2 m–2. The understory receives only 8.5 percent
distribution). For these leaf orientations, the of full sunlight. In contrast, leaves above 8.4 m
extinction coefficient is: in height receive more than 50  percent of full
sunlight.
Horizontal: Kb = 1
The exponential attenuation of light within
2 1
Vertical: Kb = (17.2) a canopy represents the mean irradiance at any
π tan Β height. This irradiance is the average of some
Spherical: K b = 0.5 sin Β
areas of the canopy in which light is unattenu-
where Β = 90o − Ζ is the elevation angle above ated and some areas that are shaded (zero irra-
the horizon and Z is solar zenith angle. diance). Eq. (17.1) also describes the fraction of
Needles tend to have a spherical leaf distri- the canopy at cumulative leaf area index x that
bution, evergreen and deciduous broadleaves is sunlit:
tend to be semi-horizontal, and grasses and
f sun ( x ) = e − K b x (17.4)
crops have semi-vertical foliage. Despite these
complexities, a common value is K b  = 0.5. Using For a canopy with leaf area index L, the sunlit
this value, a plant canopy transmits 61 percent leaf area is:
of radiation onto the ground with L = 1 m2 m–2,
37  percent with L  =  2 m2 m–2, and 22  percent Lsun = (1 − e − K b L ) K b (17.5)
with L = 3 m2 m–2. In a dense canopy with L = 6
m2 m–2, only 5 percent of solar radiation is trans- The remainder of the canopy is shaded. When
mitted through the canopy. the canopy is very dense, only a small portion is
This equation for I ↓ (z ) does not account sunlit. Most of the canopy is shaded. When the
for scattering of radiation within the canopy. Sun is high in the sky (30° zenith angle), hori-
It is derived with the assumption that leaves zontal leaves attenuate the most radiation and
completely absorb light and do not reflect or have the lowest sunlit leaf area (Figure  17.2).
transmit radiation. Scattering can be included Vertically oriented leaves have the lowest extinc-
using the equation: tion coefficient and have greater sunlit leaf area.
Maximum sunlit leaf area is only 1 m2 m–2 for
αl K b L(z )
I ↓ (z ) = I ↓ 0 e − (17.3) horizontal leaves and 1.7 m2 m–2 for spherically
oriented leaves. When the Sun is lower in the
where α l is the absorptivity of leaves (Sellers sky (60° zenith angle), much less of the canopy
1985; Campbell and Norman 1998). As absorp- is sunlit. Maximum sunlit leaf area index is less
tivity decreases, the effective extinction coef- than 1 m2 m–2 for all leaf orientations.
ficient becomes smaller and more light is The absorption, reflection, and transmit-
transmitted through the canopy. Typical absorp- tance of solar radiation by foliage have strong
tivity ranges from 0.8 to 0.2–0.3 depending on wavelength dependence. The Sun’s radiation is
solar wavelength. broadly divided into two wavebands:  the visi-
Figure 17.1 shows typical vertical profiles of ble waveband at wavelengths less than 0.7  μm
leaves and light within forest canopies. Solar and the near-infrared waveband at wavelengths
radiation decreases rapidly in the canopy. In the greater than 0.7 μm. Plants utilize light in these
oak forest, the cumulative leaf area index at a two wavebands differently (Table  17.1). Green
height of 5.8 m (70 cm into the canopy) is 1.4 m2 leaves typically absorb more than 85 percent of
m–2 and irradiance is 50 percent of full sunlight. the solar radiation in the visible waveband that
The total leaf area index is 4.6 m2 m–2, so that strikes the leaf. This light is used during photo-
the forest floor receives only about 10  percent synthesis. Light in the near-infrared waveband
of full sunlight. The aspen forest has a leaf area is not utilized during photosynthesis, and rather

Downloaded from https://www.cambridge.org/core. Danish National Library of Science and Medicine, on 25 Apr 2022 at 11:50:28, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781107339200.018
17.3  Radiative Transfer 267

(a) than absorbing this radiation, and thus possibly


1.0
overheating, leaves typically absorb less than
50 percent of the radiation in the near-infrared
0.8 waveband.
More complex models of radiative trans-
0.6 fer account for the different transmission of
Vertical
Sunlit fraction

direct beam and diffuse solar radiation and for


the scattering of radiation within the canopy
0.4
(Goudriaan 1977; Norman 1979; Sellers 1985;
Sp

Goudriaan and van Laar 1994). Figure 17.3 illus-


her
ic

trates results from the two-stream model of radi-


al

0.2

Horizontal ative transfer in plant canopies (Sellers 1985).


0.0
With low leaf area index, plants absorb little
0 1 2 3 4 5 6 7 8 solar radiation, but the absorption of radiation
Cumulative leaf area index increases substantially with higher leaf area.
(b) With a leaf area index of 4 m2 m–2, more than
3.0
90  percent of the visible radiation is absorbed.
Absorption saturates at 95 percent of the incom-
2.5 Vertical
ing radiation with a leaf area index of 6 m2
m–2. Significantly less near-infrared radiation is
Sunlit leaf area index

2.0
Spherical
absorbed. At low leaf area index, more diffuse
radiation is absorbed by the canopy than direct
1.5
beam radiation. As leaf area index increases,
Horizontal this difference becomes smaller.
1.0
The canopy consists of two types of leaves
(Figure  17.4). Sunlit leaves receive the unscat-
0.5
tered direct beam radiation absorbed by the
canopy and additionally a portion of the diffuse
0.0
0 1 2 3 4 5 6 7 8 radiation (scattered direct beam radiation and
Leaf area index atmospheric diffuse radiation) absorbed by the
Fig. 17.2  (a) Sunlit fraction and (b) sunlit leaf area index canopy. Shaded leaves receive only scattered
for horizontal, spherical, and vertical leaves in relation to leaf direct beam and atmospheric diffuse radiation.
area index with a solar zenith angle of 30°. Sunlit leaves absorb most of the direct beam
radiation; shaded leaves absorb very little direct

Table 17.1  Leaf orientation and reflection, transmission, and absorption of solar radiation by a leaf for visible
and near-infrared wavebands

Vegetation Leaf orientation Visible Near-infrared


Reflected Transmitted Absorbed Reflected Transmitted Absorbed
Needleleaf Spherical 0.07 0.05 0.88 0.35 0.10 0.55
tree
Broadleaf Semi-horizontal 0.10 0.05 0.85 0.45 0.25 0.30
tree
Grass, Semi-vertical 0.11 0.07 0.82 0.58 0.25 0.17
crop
Source: From Dorman and Sellers (1989).

Downloaded from https://www.cambridge.org/core. Danish National Library of Science and Medicine, on 25 Apr 2022 at 11:50:28, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781107339200.018
268 Plant Canopies

(a) Direct beam (b) Diffuse


1.0 1.0
Fractional canopy absorption

Fractional canopy absorption


0.8 0.8 Visible
Visible

0.6 0.6

Near-infrared Near-infrared
0.4 0.4

0.2 0.2

0.0 0.0
0 1 2 3 4 5 6 0 1 2 3 4 5 6
Leaf area index Leaf area index

Fig. 17.3  Radiative transfer in a broadleaf forest with spherical leaf orientation in relation to leaf area index. Shown for
the visible and near-infrared wavebands are (a) the fraction of direct beam solar radiation and (b) the fraction of diffuse solar
radiation absorbed by the canopy using the radiative transfer model of Sellers (1985). The zenith angle is 45° and soil albedos
are 0.10 (visible) and 0.20 (near-infrared). Leaf optical properties are from Table 17.1.

(a) Direct beam (b) Diffuse


1.0 1.0
Canopy Canopy
0.8 Sunlit 0.8
Fraction absorbed

Fraction absorbed

Sunlit
0.6 0.6

0.4 0.4 Shaded

0.2 0.2
Shaded
0.0 0.0
0 2 4 6 8 0 2 4 6 8
Leaf area index Leaf area index

Fig. 17.4  As in Figure 17.3, but showing (a) the fraction of direct beam radiation absorbed by the total canopy and the sunlit
and shaded portions of the canopy and (b) similarly for diffuse radiation. Radiative transfer uses a solution to the two-stream
approximation for sunlit and shaded leaves (Dai et al. 2004) as described by Bonan et al. (2011).

beam radiation. Diffuse radiation is more equi- albedo responds more to the optical properties
tably distributed between sunlit and shaded of foliage. Consequently, for ground surfaces
leaves. with high albedo, canopy albedo decreases with
The albedo of a plant canopy is the com- greater leaf area index. Even when the ground is
bined reflection of all plant material (leaves covered by snow, which has a very high albedo,
and stems) and the underlying ground. With plant material effectively masks the under-
low leaf area index, the albedo is largely that of lying snow. For leaf area index greater than 3
soil (Figure  17.5). As leaf area index increases, m2 m–2, the canopy albedo is similar to that of

Downloaded from https://www.cambridge.org/core. Danish National Library of Science and Medicine, on 25 Apr 2022 at 11:50:28, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781107339200.018
17.4  Canopy Photosynthesis 269

(a) Soil (b) Snow


0.30 1.0

0.25 0.8
Near-infrared
0.20
0.6
Albedo

Albedo
0.15
0.4
0.10 Near-infrared

0.05 0.2
Visible Visible
0.00 0.0
0 1 2 3 4 5 6 0 1 2 3 4 5 6
Leaf area index Leaf area index

Fig. 17.5  Canopy albedo for direct beam solar radiation in the visible and near-infrared wavebands in relation to leaf area
index. (a) Canopy albedo with a soil albedo of 0.15 (visible) and 0.30 (near-infrared). (b) Canopy albedo for snow-covered
ground with an albedo of 0.95 (visible) and 0.70 (near-infrared). Data are for a broadleaf forest (Table 17.1) with spherical leaf
orientation and for a zenith angle of 45° using the radiative transfer model of Sellers (1985).

vegetation without snow. Vegetation masking of sensing. Green leaves absorb more solar radi-
snow albedo is evident in Figure 12.10. ation in the visible waveband than in the
Surface albedo varies for direct beam near-infrared waveband (Table 17.1). In contrast,
and diffuse radiation and for the visible and other surfaces, especially soil, have smaller spec-
near-infrared wavebands. The albedo for tral differences in solar absorption. The normal-
direct beam radiation increases with greater ized difference vegetation index (NDVI), with
zenith angle. This is particularly evident for values ranging from  –1 to +1, is a measure of
spherical and semi-vertical leaf orientations. the normalized difference in reflection of solar
Consequently, albedo is higher early in the radiation in these two wavebands:
morning and late in the afternoon, when
the Sun is near the horizon, than at midday. NDVI = (rnir − rred )/(rnir + rred ) (17.6)
Figure 12.5c illustrates just such behavior in the
diurnal cycle of albedo. In contrast, the albedo where rnir and rred are reflectances in the
for diffuse radiation has no dependence on solar near-infrared and red wavebands, respectively.
zenith angle and varies only slightly with leaf Typical values for snow, lakes, and soil range
orientation. from  –0.2 to 0.05. Vegetated surfaces have val-
Despite the complexities of radiative transfer ues ranging from 0.05 to 0.70, with higher
in plant canopies, vegetation can be character- values indicating more productive vegetation.
ized by broad ranges of albedo (Table 12.1). The The NDVI is related to canopy photosynthetic
broadband albedo of vegetation averaged over capacity and allows monitoring of vegetation by
all wavelengths typically ranges from 0.05 to remote sensing (Tucker et al. 1985, 1986, 1991;
0.25. Forests generally have lower albedo than Myneni et al. 1997).
grasslands or croplands. Coniferous forests have
lower albedo than deciduous forests.
The different optical properties of foliage in 17.4  Canopy Photosynthesis
the visible and near-infrared wavebands is an
important and distinguishing spectral charac- The photosynthetic uptake of carbon by a can-
teristic of vegetation that facilitates the mon- opy of leaves (i.e., gross primary production;
itoring and analysis of vegetation by remote GPP) is the sum of the photosynthetic rates of the

Downloaded from https://www.cambridge.org/core. Danish National Library of Science and Medicine, on 25 Apr 2022 at 11:50:28, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781107339200.018
270 Plant Canopies

(a) Sorghum (b) Tobacco Fig. 17.6 Vertical profile of


stomatal conductance in crops.
75 150
1600 (a) Sorghum at 0700, 1200, and
1830 1600 hours. (b) Tobacco at 0830,
0700 1200 1200, and 1830 hours. Adapted
0830
from Jarvis and McNaughton (1986).
50 100
Height (cm)

Height (cm)
1200

25 50

0 0
0 0.1 0.2 0 0.2 0.4
Stomatal conductance Stomatal conductance
(mol H2O m–2 s–1) (mol H2O m–2 s–1)

individual leaves, accounting for within-canopy day, stomatal conductance is greater at the top
variations in light, foliage nitrogen, and other of the canopy than deeper in the canopy where
factors. One approach to estimate canopy pho- leaves are shaded. The total carbon uptake dur-
tosynthesis is based on the observation that ing photosynthesis requires integration of leaf
biomass accumulation is proportional to the photosynthesis over the light profile.
amount of radiation intercepted by the canopy Norman (1993) reviewed different meth-
(Monteith 1977). Production efficiency models odologies for this leaf-to-canopy scaling. One
simulate carbon assimilation proportional to method is to assume that all leaves in the can-
absorbed photosynthetically active radiation opy receive the average radiation absorbed by
times a light-use efficiency, which relates carbon the canopy, but this fails to account for the
gain to light absorbed (Prince and Goward 1995; non-linear extinction of light within the canopy
Running et al. 2000, 2004; Zhao et al. 2005; Yuan and the non-linear dependence of photosynthe-
et al. 2007). A typical light-use efficiency is 1.5 g sis and stomatal conductance on light. Scaling
C per MJ. This potential productivity is reduced techniques need to account for the decrease in
by environmental constraints as: light with greater depth in the canopy. In par-
ticular, some leaves are sunlit and receive full
GPP = Ε F ↓ f1 (T ) f 2 (θ) f3 ( D) (17.7)
illumination. Others are shaded and have low
where Ε is light-use efficiency, F ↓ is absorbed rates of photosynthesis.
photosynthetically active radiation, and f1 (T ), One scaling approach analytically integrates
f 2 (θ), and f3 ( D) are empirical functions scaled leaf photosynthesis over the light profile using a
from zero to one that adjust photosynthesis simple model of the photosynthetic response to
for temperature, soil water, and vapor pressure light combined with exponential attenuation of
deficit. light in the canopy (Sellers 1985; Norman 1993).
Other methods explicitly scale leaf fluxes to The increase in the rate of leaf photosynthesis
the canopy by integrating photosynthesis and with more absorbed photosynthetically active
stomatal conductance over the light profile. radiation can be represented by:
Figure 17.6 illustrates variation of stomatal con- A max ΕF ↓
ductance with height in two canopies of crop Al = (17.8)
ΕF ↓ + A max
plants. Stomatal conductance varies little with
height early in the morning when low light lev- where A max is the maximum rate of photosyn-
els limit the rate of photosynthesis. Later in the thesis at saturating light (μmol m–2 s–1), F ↓ is

Downloaded from https://www.cambridge.org/core. Danish National Library of Science and Medicine, on 25 Apr 2022 at 11:50:28, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781107339200.018
17.4  Canopy Photosynthesis 271

(a) 20 (b) 60

50 F0 = 1000
(µmol CO2 m–2 s–1)
15

Canopy assimilation
(µmol CO2 m–2 s–1)
Leaf assimilation

40 F0 = 500

10 30
F0 = 250
20
5
10

0 0
0 500 1000 1500 2000 0 2 4 6 8 10
APAR (µmol m–2 s–1) Leaf area index

Fig. 17.7  Canopy scaling of photosynthesis. (a) Leaf photosynthesis in relation to absorbed photosynthetically active radiation
(APAR). (b) Canopy photosynthesis in relation to leaf area index obtained by integrating leaf photosynthesis over an exponential
light profile. Canopy photosynthesis is shown for three values of incident PAR at the top of the canopy.

absorbed photosynthetically active radiation receive diffuse radiation and additionally the
(μmol m–2 s–1), and Ε is light-use efficiency (mol direct beam radiation intercepted in the canopy.
mol–1). If light is attenuated through the canopy Because they receive much more radiation than
exponentially as in Eq. (17.1), the integral of A l shaded leaves, sunlit leaves are typically near
with respect to leaf area index (L) is the total light saturation. Canopy photosynthesis is the
photosynthesis of the canopy, equal to: sum of these two rates, each multiplied by their
respective leaf area index:
A max  b + F0 
GPP = ln   (17.9)
Kb  b + F0 exp(− K b L)  GPP = A sun Lsun + A shade Lshade (17.10)

where b = A max / Ε and F0 is the photosyntheti- This approach has the advantage of accounting
cally active radiation incident on the canopy. for the different attenuation of direct beam and
Figure 17.7 shows this scaling for A max  = 20 μmol diffuse radiation within the canopy.
m–2 s–1, Ε = 0.06 mol mol–1, and K b  = 0.5. In this Leaves growing in shade generally have
example, leaf photosynthesis attains a rate of lower photosynthetic capacity than leaves grow-
17 μmol m–2 s–1 at high irradiance. Canopy pho- ing in sun (Table  16.1). This is a result of dif-
tosynthesis increases with greater leaf area ferent investments in chlorophyll and Rubisco.
index as more photosynthetically active radia- These investments require much nitrogen and
tion is absorbed by foliage, but saturates with are energetically expensive to maintain. Plants
leaf area index of 6–7 m2 m–2 as light absorption that invest too much in photosynthetic capac-
also saturates. ity will be at a competitive disadvantage if this
Another method to integrate photosynthesis machinery is underutilized, such as occurs in
over the canopy is to consider the fraction of the a shaded understory. Plants that match photo-
canopy that is sunlit and shaded (Norman 1993). synthetic capacity to local resource availability
Different leaf photosynthetic rates are calcu- have an advantage. Based on this reasoning,
lated for these two classes of leaves based on the the vertical profile of leaf nitrogen and photo-
amount of absorbed photosynthetically active synthetic capacity through the canopy should
radiation. Shaded leaves receive only diffuse be related to the light profile. This concept
radiation, from both the sky and direct beam provides another means, similar to sunlit and
radiation scattered within the canopy, and are shaded leaves, to integrate photosynthesis and
typically in the linear portion of the photo- stomatal conductance for a canopy (Sellers
synthetic light response curve. Sunlit leaves et al. 1992).

Downloaded from https://www.cambridge.org/core. Danish National Library of Science and Medicine, on 25 Apr 2022 at 11:50:28, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781107339200.018
272 Plant Canopies

(a) (b) (c)


12
Photosynthetically active radiation (%)

Photosynthetic capacity (µmol m–2 s–1)


100 2.0

10
80

Foliage nitrogen (g m–2)


1.5 8
60
6
40
1.0 4

20 2

0 0.5 0
0 2 4 6 8 10 12 14 16 0 20 40 60 80 100 0.5 1.0 1.5 2.0
Canopy depth (m) Photosynthetically active radiation (%) Foliage nitrogen (g m–2)

(d) (e) (f)


100 100 20
Relative photosynthetic capacity (%)

Relative photosynthetic capacity (%)

Canopy photosynthetic capacity


Aspen
15 Black
90
90

(µmol m–2 s–1)


spruce
10
80 Young jack pine Black
80 Old jack pine spruce
Old black spruce 5
70 Old jack
Old aspen Young pine
70 0 jack pine
70 80 90 100 0 20 40 60 80 100 0.5 0.6 0.7 0.8 0.9
Relative nitrogen content (%) Photosynthetically active radiation (%) Normalized difference vegetation index

Fig. 17.8  Canopy integration of photosynthetic capacity based on relationships among light, foliage nitrogen, and
photosynthetic capacity for jack pine, black spruce, and quaking aspen forests. (a) Photosynthetically active radiation as a
percentage of full radiation (%PAR) in relation to depth from the top of the canopy. (b) Foliage nitrogen in relation to %PAR.
(c) Leaf photosynthetic capacity in relation to foliage nitrogen. (d) Leaf photosynthetic capacity (as a percentage of that in the
upper canopy) in relation to leaf nitrogen (as a percentage of that in the upper canopy). (e) Leaf relative photosynthetic capacity
in relation to %PAR. (f) Canopy photosynthetic capacity integrated over all leaves in relation to the normalized difference
vegetation index. Data from Dang et al. (1997).

Data collected in the boreal forests of foliage high in the canopy receives more radi-
Canada illustrate how light, foliage nitrogen, ation, has greater nitrogen concentration,
and photosynthetic capacity within tree cano- and has greater photosynthetic capacity than
pies are interrelated (Figure 17.8). In jack pine, foliage lower in the canopy. Indeed, through-
black spruce, and aspen forests, photosynthet- out the canopy there is a linear relationship
ically active radiation decreases exponentially between photosynthetic capacity at a given
with greater depth from the top of the canopy. height and the corresponding nitrogen con-
Foliage nitrogen decreases with decreasing tent at the same height, both expressed rela-
photosynthetically active radiation. Maximum tive to that at the top of the canopy. Relative
photosynthetic rates under light-saturated photosynthetic capacity decreases exponen-
conditions increase with greater foliage nitro- tially with decreasing photosynthetically
gen. Jack pine and black spruce trees differ active radiation.
little in this relationship; aspen trees have The original theory postulated that plants
a higher photosynthetic capacity for a given optimally allocate resources to maximize car-
nitrogen concentration. These data show that bon gain such that area-based leaf nitrogen is

Downloaded from https://www.cambridge.org/core. Danish National Library of Science and Medicine, on 25 Apr 2022 at 11:50:28, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781107339200.018
17.5  Canopy Conductance 273

distributed through the canopy in relation to


the time-mean profile of photosynthetically
17.5  Canopy Conductance
active radiation, but it is now recognized that
the nitrogen gradient is shallower than the At the scale of an individual leaf, stomatal con-
light gradient (Hollinger 1996; Carswell et  al. trol of transpiration is quantified by stomatal
2000; Meir et  al. 2002; Niinemets 2007; Lloyd conductance. At the scale of a canopy of leaves,
et  al. 2010). The decline in leaf nitrogen and an aggregate measure of surface conductance is
photosynthetic capacity is described by an expo- required to account for evaporation from soil
nential relationship with cumulative leaf area and transpiration from foliage. In particular, the
index, similar to the sunlit fraction: physiology of individual leaves, which is directly
measurable in terms of the response of stomata
f n ( x ) = e − K n x (17.11) to light, water, temperature, and other environ-
mental factors, must be scaled over all leaves
Lloyd et al.’s (2010) estimates of K n for 16 temper- to the canopy, where photosynthesis and tran-
ate broadleaf forests and 2 tropical forests range spiration can only be measured for the aggre-
from 0.10 to 0.43 (mean, 0.19; median, 0.18). gate canopy. Whereas stomatal conductance
The parameters Vc max and J max in the describes the conductance of an individual leaf,
Farquhar et  al. (1980) photosynthetic model canopy conductance describes the aggregate
vary with leaf nitrogen (Rogers 2014). Many conductance of a canopy of leaves. Surface con-
plant canopy models parameterize canopy ductance additionally includes soil evaporation.
scaling using concepts of sunlit and shaded In dense canopies, soil evaporation is negligible
leaves in combination with an exponential pro- and surface and canopy conductances are effec-
file of foliage nitrogen (de Pury and Farquhar tively the same. Canopy or surface conductance
1997; Wang and Leuning 1998; Dai et al. 2004; can be derived from micrometeorological mea-
Bonan et  al. 2011). The canopy is divided surements of fluxes above the canopy.
into sunlit and shaded fractions, and the Monteith (1965) extended the Penman
photosynthesis–conductance model is solved equation to a canopy of leaves, and the
using canopy-integrated parameters derived Penman–Monteith equation is widely used for
from leaf-level parameters. Canopy values for biometeorological studies. Eq. (12.22) is the
Vc max are found by integrating leaf nitrogen Penman–Monteith equation for a canopy, and:
concentration, using Eq. (17.11), over the sun-
lit and shaded fractions of the canopy, using s ( R n − G ) + c p e (Ta ) − ea  g ah
λE =  *  (17.14)
Eq. (17.4). Other parameters scale similarly. s + γ (1 + g ah / g c )
For example, Vc max integrated over the sunlit
canopy is: This equation shows that the prominent envi-
ronmental controls on evapotranspiration are
L available energy (R n − G ), atmospheric vapor
Vc max (sun ) = ∫ Vc max 0 f n ( x ) f sun ( x ) dx pressure deficit (e* (Ta ) − ea), and aerodynamic
0 (17.12) (g ah ) and canopy (g c ) conductances.
1
= Vc max 0 1 − e −( K n + K b ) L  The Penman–Monteith equation is a bulk
Kn + Kb
surface formulation in which two conductances
acting in series regulate latent heat exchange
where Vc max 0 is the value at the top of the can-
with the atmosphere. The canopy conductance
opy. For the shaded canopy:
(g c ) governs processes within the plant canopy,
L and the aerodynamic conductance (g ah ) governs
Vc max (sha) = ∫ Vc max 0 f n ( x ) 1 − f sun ( x ) dx turbulent processes above the canopy. As with
0   (17.13) an individual leaf, the canopy has degrees of
1
= Vc max 0 [1 − e − Kn L
] K − Vc max (sun) coupling to the atmosphere determined by the
n magnitude of the aerodynamic conductance

Downloaded from https://www.cambridge.org/core. Danish National Library of Science and Medicine, on 25 Apr 2022 at 11:50:28, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781107339200.018
274 Plant Canopies

(a) (b) 1.7 (c) 1.7


gsmax= 0.5

4 LAI = 10

Gsmax (mol H2O m-2 s-1)

Gsmax (mol H2O m-2 s-1)


1.3 1.3
Gs / gsmax
3 gsmax= 0.3
Ratio

0.8 0.8
Gc / gsmax LAI = 2
2

0.4 gsmax= 0.2 0.4


1
Gs / Gc Gsmax = 2.7 gsmax

0 0 0
0 2 4 6 8 10 0 2 4 6 8 10 12 0 0.17 0.34 0.51
Leaf area index Leaf area index gsmax (mol H2O m-2 s-1)

Fig. 17.9  Relationships among maximum leaf, canopy, and surface conductance when vegetation is not stressed by soil
moisture and other environmental conditions. (a) Theoretical relationships among maximum leaf conductance (g s max ), canopy
conductance (Gc ), and surface conductance (Gs ) with increasing leaf area index for g s max  = 0.3 mol m–2 s–1. Relationships are
for particular values of net radiation, vapor pressure deficit, aerodynamic conductance, and response of stomata to light.
(b) Observed maximum surface conductance (Gs max ) in relation to leaf area index for a variety of grasslands, forests, and
croplands. The solid lines show the expected relationship for three values of g s max . (c) Observed Gs max in relation to g s max for
several grasslands, forests, and croplands. The dashed lines show the lower and upper limits of Gs max expected from theory with
leaf area index of 2 and 10 m2 m–2. Adapted from Kelliher et al. (1995).

(Jarvis and McNaughton 1986). With a low aero- The Penman–Monteith equation gives
dynamic conductance, the canopy is decoupled insight to surface and canopy conductances and
from the atmosphere, and evapotranspiration their relationship with leaf area index and leaf
approaches an equilibrium rate dependent on conductance (Schulze et al. 1994; Kelliher et al.
available energy: 1995). The integrated source strength of a can-
opy is proportional to leaf conductance times
λ E = s ( R n − G ) ( s + γ ) (17.15)
leaf area, and canopy conductance increases
The canopy has no influence on the rate, and linearly with leaf area index at low leaf area
the rate represents a freely evaporating surface. (Figure  17.9a). Canopy conductance is rela-
Where there is a high aerodynamic conductance, tively constant at a value largely determined
evapotranspiration approaches a rate imposed by by leaf conductance at high leaf area. Canopy
vapor pressure deficit and canopy conductance: conductance becomes constant with high leaf
area because stomata close due to low light
cp
λE = e* (Ta ) − ea  g c (17.16) levels deep in the canopy. Surface conductance
γ  approaches canopy conductance as leaf area
increases because soil evaporation becomes neg-
The Penman–Monteith equation can be
ligible in a dense canopy. Surface conductance
re-arranged to give the canopy conductance:
significantly exceeds canopy conductance only
1 ε + 1  ε ( Rn − G )  c p e* (Ta ) − ea  for low leaf area index less than about 3 m2 m–2,
=  − 1 + (17.17) where soil evaporation is more important. For a
gc g ah  ( ε + 1) λ E  γ λE
given leaf conductance, surface conductance is
with ε = s / γ . By measuring evapotranspiration nearly independent of leaf area because the ten-
from a canopy and if all other terms are known, dency for canopy conductance to decrease with
the Penman–Monteith equation can be solved low leaf area is countered by increasing propor-
for g c . This conductance captures the effects tion of soil evaporation.
of leaf area, canopy coverage, photosynthetic Across a variety of forests, grasslands, and
capacity, and soil moisture on the partitioning croplands, surface and canopy conductances
of net radiation into latent heat flux. derived from the Penman–Monteith equation

Downloaded from https://www.cambridge.org/core. Danish National Library of Science and Medicine, on 25 Apr 2022 at 11:50:28, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781107339200.018
17.5  Canopy Conductance 275

Table  17.2  Maximum leaf stomatal conductance and bulk canopy conductance achieved in the field for
unstressed, well-lit leaves

Plant group Leaf (mol H2O m–2 s–1) Canopy (mol H2O m–2 s–1)

Woody plants 0.23 0.83


Natural herbaceous plants 0.34 0.72
Agricultural crops 0.49 1.34
Note: Converted to mol m–2 s–1 using, ρm = 42.3 mol m–3.
Source: From Kelliher et al. (1995) and Körner (1994).

2.0 Fig. 17.10  Latent heat flux


normalized by its equilibrium rate
Wheat
in relation to canopy resistance for
Corn
wheat, corn, temperate deciduous
Deciduous forest
1.5 forest, boreal jack pine conifer forest,
Jack pine
Oak savanna and oak savanna. Shown are individual
Normalized latent heat flux

data points and the mean for each


vegetation type. Original data from
Baldocchi et al. (1997) and Baldocchi
1.0 and Xu (2007).

0.5

0.0
10 100 1000 10,000
Canopy resistance (s m ) –1

show relationships similar to that expected for unstressed, well-lit leaves in three types
from theory. Maximum surface conductance of plants. A  leaf conductance of 0.23 mol H2O
in relation to leaf area index is within the m–2 s–1 is typical for woody plants. Herbaceous
bounds determined by three values of leaf con- plants have a higher conductance, and crops
ductance representative of woody plants, nat- have the highest leaf conductance. Canopy con-
ural herbaceous plants, and agricultural crops ductance is higher because of the high leaf area
(Figure  17.9b). Maximum surface conductance index of the various plant canopies. Greater leaf
increases linearly with maximum leaf conduc- area index increases latent heat exchange with
tance with a slope of ~3 (Figure  17.9c). This is the atmosphere by increasing the surface area
consistent with theory, which predicts that for from which moisture is lost.
any leaf conductance surface conductance is Similar differences are seen among other
constrained within limits set by leaf area and vegetation types. Figure  17.10 compares the
that surface conductance is largely determined rate of latent heat flux normalized by the equi-
by leaf conductance at high leaf area. librium rate in relation to canopy resistance
Table  17.2 compares maximum leaf sto- (the inverse of canopy conductance) for sev-
matal conductance and canopy conductance eral types of vegetation. Equilibrium evapo-
derived from the Penman–Monteith equation transpiration, given by Eq. (17.15), is that for

Downloaded from https://www.cambridge.org/core. Danish National Library of Science and Medicine, on 25 Apr 2022 at 11:50:28, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781107339200.018
276 Plant Canopies

a well-watered site and depends on available models that follow the trajectories of particles
energy. The ratio of latent heat flux to the equi- (Baldocchi 1992).
librium flux is a measure of the partitioning Studies of turbulence in a deciduous forest
of available energy to evapotranspiration. The illustrate some of the challenges in describing
evaporative fraction is generally high in agri- turbulent transfer in plant canopies (Baldocchi
cultural crops, and these plants have low evap- and Meyers 1988a,b, 1989; Meyers and Baldocchi
orative resistance (high canopy conductance). 1991). The particular forest studied consists of
Forests have lower evaporative fractions com- oak and hickory trees with a canopy height of
pared with crops and higher resistance (lower about 23 m and a leaf area index of about 5 m2
conductance). Dry jack pine forest and oak m–2. Most leaf area is in the upper 5 m of the
savanna have the lowest evaporative fraction canopy. Over 75 percent of the total leaf area is
and highest resistance. located in the upper 25 percent of the canopy.
Measurements in the canopy reveal a complex
17.6 Turbulent Transfer in Forest turbulent structure that is highly intermittent
and characterized by non-Gaussian probability
Canopies distributions. Wind speed within the canopy var-
ies with height in relation to the profile of leaf
Many models of turbulent fluxes in plant can- area index (Figure 17.11). Strong shear occurs in
opies utilize the principle of diffusion along the upper 20 percent of the canopy above 0.8h,
the mean concentration gradient, similar to where much of the leaf area is concentrated. In
fluxes in the surface layer (Chapter  13). The this region, wind speed decreases sharply with
fluxes of momentum, sensible heat, and water depth in the canopy. A  reversal in the vertical
vapor are assumed to be proportional to the gradient of wind speed occurs below this height
vertical gradients of wind, temperature, and with a local maximum in the canopy at about
specific humidity, respectively, multiplied by a 0.5h. Wind speed thereafter decreases towards
turbulent diffusivity, or aerodynamic conduc- the forest floor. This wind profile is not a simple
tance. However, Monin–Obukhov stability the- exponential profile as commonly assumed, and
ory fails in the layer of air immediately above illustrates counter-gradient momentum trans-
the canopy, known as the roughness sublayer, port in the canopy.
and the universal similarity functions are not Denmead and Bradley (1985) found exam-
valid immediately above or within plant cano- ples of counter-gradient or zero-gradient
pies. The gradient–diffusion approach also fails fluxes in their measurements in a pine forest.
within plant canopies, where counter-gradient Figure 17.12 shows profiles of potential temper-
or zero-gradient fluxes and intermittent tur- ature, water vapor mixing ratio, and CO2 concen-
bulence are common (Denmead and Bradley tration measured in the canopy. Temperature
1985; Raupach and Finnigan 1988; Baldocchi has a maximum near the middle of the canopy.
1989; Baldocchi and Meyers 1998; Finnigan Water vapor decreases from a high near the
2000). More detailed models of fluid dynamics surface, is nearly constant with height in the
than gradient–diffusion theory are needed to lower canopy space, and decreases with height
adequately represent turbulence in plant cano- in the upper canopy. The concentration of CO2
pies. Such models include higher-order closure has a minimum in the middle of the canopy.
models based on equations for the first-order From these gradients, one would expect upward
moment (e.g., mean horizontal wind velocity, sensible heat flux above the temperature max-
mean mixing ratio) and second-order moments imum and downward sensible heat flux onto
(e.g., vertical velocity variance, covariance of the forest floor below mid-canopy. The steep
mixing ratio and vertical velocity fluctuations) gradient of water vapor near the forest floor
associated with turbulent transfer (Wilson suggests enhanced evaporation from the forest
and Shaw 1977; Meyers and Paw U 1986, 1987; floor. The CO2 profile implies upward transport
Wilson 1988; Pyles et  al. 2000) or Lagrangian of CO2 from the forest floor in the lower canopy

Downloaded from https://www.cambridge.org/core. Danish National Library of Science and Medicine, on 25 Apr 2022 at 11:50:28, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781107339200.018
17.6 Turbulent Transfer in Forest Canopies 277

(b) 1.4

1.2

(a) 1.0 1.0

0.8 0.8

Height (z / h)
Height (z / h)

0.6 0.6

0.4 0.4

0.2 0.2

0 0
0 0.2 0.4 0.6 0.8 1.0 0 1 2 3 4 5 6 7
Plant area density (m2 m-3) u / u*

Fig. 17.11  Profiles of (a) leaf area and (b) wind speed above and in the canopy of a deciduous forest. Height (z) is given as
a fraction of canopy height (h). Wind speed is normalized by friction velocity (u* ) measured above the canopy. Adapted from
Baldocchi and Meyers (1988a) and Baldocchi (1989).

(a) Temperature (b) Mixing ratio (c) CO2 Fig. 17.12  Mean profiles
of (a) potential temperature,
(b) water vapor mixing ratio,
Fc = –0.54
and (c) CO2 concentration
observed in a ponderosa pine
20 forest over a period of one
hour. Also shown are the
Height (m)

H = 433 λE = 296
Fc = –0.23 fluxes of sensible heat
(H, W m–2), latent heat
(λE, W m–2), and CO2
(Fc , mg m–2 s–1) at two heights
6 in the canopy. Adapted from
Denmead and Bradley (1985).
H = 64 λE = 59

29 30 31 10.0 10.2 10.4 327 328 329


°C g kg–1 ppm

and downward CO2 flux in the upper canopy. water vapor. The downward flux of CO2 occurs
Observed fluxes show the opposite. Fluxes in along a positive concentration gradient. In the
the upper canopy conform to the conventional lower canopy, however, these fluxes are associ-
gradient–diffusion relationships. Upward fluxes ated with counter-gradients or zero-gradients.
of sensible heat and water vapor are accompa- The sensible heat flux is upward even though
nied by negative gradients of temperature and temperature increases with height. A  large

Downloaded from https://www.cambridge.org/core. Danish National Library of Science and Medicine, on 25 Apr 2022 at 11:50:28, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781107339200.018
278 Plant Canopies

Sensible heat flux

Ta Ta Ta

gah
gah
gah
Tac Tv
Ts gch
gac
Ts Tg

Latent heat flux

ea ea ea

gaw
gaw
gaw gsw
gcw ev
es eac
gsw gac
es eg

(a) Bulk surface (b) Bulk surface (c) Two-source (d) Multi-layer canopy
without canopy with canopy canopy

Fig. 17.13  Conductance networks for sensible heat flux (top) and latent heat flux (bottom). Networks are for a bulk surface
without a canopy, vegetation with a bulk surface formulation, a two-source canopy, and multiple canopy layers. (a) The surface is
the ground and fluxes are regulated by aerodynamic conductances (g ah , g aw ). Latent heat flux also includes a soil conductance
(g sw ). (b) The bulk canopy uses an effective surface temperature (Ts ) and vapor pressure (es) that combines vegetation and
ground. The overall conductance for latent heat is a surface conductance (g sw ) acting in series with an aerodynamic conductance
(g aw ). (c) The two-source canopy partitions fluxes into vegetation and ground based on vegetation (Tv,ev ) and ground
(Tg ,e g ) temperature and vapor pressure. The conductance g ac accounts for aerodynamic processes within the canopy. Canopy
conductances account for sensible heat exchange (g ch) from the vegetation to air within the canopy (Tac ) and latent heat
exchange (g cw ) from the vegetation to canopy air (eac ). For latent heat, this also includes stomatal conductance. (d) Multiple
canopy layers.

upward flux of water vapor occurs despite zero a single aerodynamic conductance between
gradient. Substantial absorption of CO2 occurs the surface and atmosphere, as described in
below the middle of the canopy. Chapters  12 and 13. A  soil conductance can
account for the extent to which the surface is
not saturated with moisture.
17.7 Canopy Models The effects of vegetation on surface fluxes can
be included by treating the soil–canopy system
The same principles that determine the temper- as an effective bulk surface, analogous to that
ature and energy fluxes of a leaf also determine for a non-vegetated surface but with radiative
the temperature and energy fluxes from vege- exchange integrated over the canopy and addi-
tated surfaces. Figure  17.13 illustrates some of tional conductances for leaves (Figure  17.13b).
the ways in which turbulent fluxes are modeled. This type of formulation is referred to as a
One approach is to describe these fluxes simply big-leaf model because the canopy is treated as a
through a bulk aerodynamic formulation with single leaf scaled to represent a canopy. Canopy

Downloaded from https://www.cambridge.org/core. Danish National Library of Science and Medicine, on 25 Apr 2022 at 11:50:28, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781107339200.018
17.7 Canopy Models 279

flux equations in a big-leaf model are similar to H = 200 W m–2, and an air temperature Ta= 30°C.
those of an individual leaf, but with a leaf con- The temperature at z 0 h is found by evaluating
ductance representative for all the leaves in the the bulk aerodynamic formula for sensible heat
canopy. The Penman–Monteith equation is a with g ah , using Eq. (12.10):
big-leaf model for latent heat flux.
Bulk surface models can be used to monitor Ts = Ta + H c p g ah (17.19)
plant canopies. If sensible heat flux and aerody-
namic conductance are known from micromete- so that Ts = 34.6°C (with c p = 29.2 J mol–1 K–1).
orological measurements, the effective surface Similarly, the temperature at z 0 m , evaluated
temperature (Ts ) is estimated from the bulk aero- using g am rather than g ah , is Ts = 32.2°C. The tem-
dynamic formulation for sensible heat using perature difference between Ts evaluated at z 0 m
Eq. (12.10). In this case, Ts represents the aero- and Ts evaluated at z 0 h is:
dynamic temperature of the canopy, regulated
Ts ( z 0 m ) − Ts ( z0 h ) 1  z0 m 
by the aerodynamic conductance between the = ln  (17.20)
θ* k  z0 h 
atmosphere and the canopy. Alternatively, esti-
mates of Ts can be inferred from measurement where θ* = −H / (ρm c p u* ). In this example,
of upward longwave radiation using Eq. (12.8). θ* = –0.42 K (with ρm = 42.3 mol m–3).
Here, Ts represents the temperature at which By influencing aerodynamic conductance,
the effective surface emits longwave radiation the height of plants has a large influence on
and is known as the radiometric temperature. leaf temperature. Tall forest vegetation is aero-
In applying bulk aerodynamic formula, it is dynamically rough and has a high aerody-
necessary to distinguish the roughness length namic conductance (Figure 13.6). Heat is readily
for momentum (z 0 m ) from the roughness length exchanged with the atmosphere, and the tem-
for heat (z 0 h), as discussed in Chapter  13. The perature of leaves is closely coupled to that of
latter is typically taken to be less than that of the air. In contrast, short vegetation such as
momentum (Thom 1972; Garratt and Hicks grass or shrub is aerodynamically smooth, has
1973; Garratt 1978; Beljaars and Holtslag 1991). a low aerodynamic conductance, and dissipates
If not, there can be significant difference in the heat less effectively. The leaves of short vegeta-
surface temperature inferred from the sensible tion are decoupled from the air and have warmer
heat flux. Consider, for example, a forest canopy temperatures than that of the air. In cold cli-
with a height of 20 m, displacement height d = mates, short stature may convey an advantage
14 m, and roughness lengths z 0 m = 2 m and z 0 h = by warming leaf temperature (Wilson et al. 1987;
0.2 m. The wind speed at height z= 30 m is u= 2 Grace 1988; Grace et al. 1989). Figure 17.14 illus-
m s–1, and for convenience the effects of atmo- trates this for alpine forest and shrub vegeta-
spheric stability are neglected. From Eq. (13.26), tion. For both types of vegetation, the leaf-to-air
the aerodynamic conductances evaluated using temperature difference increases with greater
z 0 m and z 0 h are g am= 3.1 and g ah = 1.5 mol m–2 s–1. net radiation. In the tall forest, the slope of this
The decreased conductance (i.e., the excess relationship is small. Leaf temperatures are sim-
resistance) between z 0 m and z 0 h is 2.8 mol m–2 s–1 ilar to air temperature, with a maximum excess
(1 / g b* = 1 / g ah − 1 / g am). A  general expression of less than 5°C. In the dwarf shrub vegetation,
for the excess resistance between heights z 0 m the slope is larger and the temperature excess is
and z 0 h is: 15°C in bright sunshine.
1 ln ( z 0 m z 0 h ) As an alternative to the bulk representation
= (17.18) of the effective surface for heat and moisture
g b* ρm ku *
exchange, the surface can be explicitly repre-
In this example, the friction velocity is u * = 0.385 sented by ground and vegetation (Figure 17.13c).
m s–1, from Eq. (13.21). This excess resistance Sensible heat is partitioned into that from
translates into a higher temperature at z 0 h than foliage and that from soil, each regulated by
at z 0 m . For example, assume a sensible heat flux different processes. The leaf boundary layer

Downloaded from https://www.cambridge.org/core. Danish National Library of Science and Medicine, on 25 Apr 2022 at 11:50:28, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781107339200.018
280 Plant Canopies

20 Here, g ah is the aerodynamic conductance for


Leaf–air temperature difference (oC) sensible heat, g ch is the leaf boundary layer
15 conductance scaled to the canopy, and g ac is an
aerodynamic conductance within the canopy.
10 Shrub
Rearranging terms in these equations, the can-
opy air temperature (Tac ), which is common to
5 all three fluxes, is evaluated as a weighted aver-
age of the atmospheric, vegetation, and ground
0 Forest
temperatures (Ta, Tv, and Tg ):

–5 g ahTa + g chTv + g acTg


0 200 400 600 800 1000 Tac = (17.22)
g ah + g ch + g ac
Net radiation (W m-2)

Fig. 17.14  Relationship between daytime leaf–air This equation for canopy air temperature is
temperature difference and net radiation for forest substituted into the expression for H v. A similar
(15–18 m tall) and dwarf shrub (0.1 m tall). Data are shown equation is derived for canopy air vapor pres-
for wind speeds of 1–3 m s–1. Data from Grace et al. (1989). sure (eac ) for latent heat flux, and the energy bal-
ance of the canopy is solved for the temperature
(Tv) that balances net radiation, sensible heat
conductance integrated over all leaves in the
flux, and latent heat flux. Then, with canopy
canopy governs sensible heat flux from foliage.
fluxes known, the ground fluxes and tempera-
Turbulent processes within the canopy govern
ture (Tg ) are updated.
sensible heat flux from the soil. Latent heat is
Alternatively, one can assume that the can-
partitioned into soil evaporation and transpira-
opy air has some capacity to store heat (Vidale
tion. Transpiration is regulated by a canopy con-
and Stöckli 2005). The storage of heat in the can-
ductance that is an integration of leaf boundary
opy air is:
layer and stomatal conductances over all the
leaves in the canopy. Soil evaporation is regu- ρm c p ( ∂T / ∂t ) ∂z (17.23)
lated by aerodynamic processes within the plant
canopy and by soil moisture. These equations can where ∂T / ∂t is the rate of change of tempera-
be solved by writing the sensible and latent heat ture (K s–1) and ∂z is canopy height (m). For a
fluxes as a linear combination of atmospheric, canopy with a height z = 25 m, a change in tem-
vegetation, and ground temperatures and vapor perature of 1°C hr–1 is associated with approx-
pressure, respectively (Deardorff 1978). imately 8 W m–2 heat storage in the canopy. If
The leaf and ground energy budgets repre- heat storage is included, the canopy air temper-
sent a system of equations that can be solved ature is not diagnosed as a linear combination of
for the unknown vegetation (Tv) and ground Ta, Tv, and Tg as above, but rather predicted from:
(Tg ) temperatures. With reference to the
one-layer canopy depicted in Figure  17.13c, ρm c p ∆z ( ∆Tac ∆t ) = − H + H v + H g (17.24)
three sensible heat fluxes are represented in
the soil–plant–atmosphere system:  the flux The storage heat in the canopy air space is the
from vegetation to canopy air (H v); from ground difference between heat entering the air space
to canopy air (H g ); and from canopy air to the (H v,H g ) and sensible heat transferred to the
atmosphere (H). Assuming the canopy air has atmosphere (H).
negligible capacity to store heat, the total sen- The preceding methodologies provide
sible heat flux to the atmosphere is the sum of simple means to simulate fluxes from plant
fluxes from vegetation and the ground: canopies and are commonly used in climate
models (Deardorff 1978; Dickinson et  al. 1986,
H = c p (Tac − Ta ) g ah = c p (Tv − Tac ) g ch + c p (Tg − Tac ) g ac 1993; Sellers et  al. 1986, 1996a,b). They can be
 (17.21) extended to represent multiple canopy layers

Downloaded from https://www.cambridge.org/core. Danish National Library of Science and Medicine, on 25 Apr 2022 at 11:50:28, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781107339200.018
17.8 Environmental Controls of Canopy Fluxes 281

(a) Leafless Fig. 17.15 Average


CO2 concentration (ppm) diurnal cycle of CO2
370 concentration within and
above a 70-year-old quaking
368 aspen forest with a canopy
0.8 m
366 height of 21.5 m when
2.3 m (a) leafless and (b) in full
364 leaf. Adapted from Yang et al.
≥ 9.5 m (1999).
362

(b) Full leaf


550
CO2 concentration (ppm)

500
0.8 m

450
2.3 m

400
≥ 9.5 m
350

0000 0400 0800 1200 1600 2000 2400


Local time (hours)

(Figure 17.13d) with a simple representation of has a negative radiative balance. Sensible and
aerodynamic conductances within the canopy latent heat fluxes are small. Stomata are closed,
(Shuttleworth and Wallace 1985; Choudhury and trees do not absorb CO2 during photosyn-
and Monteith 1988). However, two-source can- thesis. Carbon dioxide is, however, lost during
opy models and multilayer models are based respiration, and there is a net flux of CO2 from
on gradient–diffusion theory and do not allow the forest to the atmosphere. As the Sun rises
for counter-gradient fluxes as observed in plant and solar radiation is absorbed by the forest,
canopies. More complex canopy models are there is a net gain of radiation at the surface.
required to resolve turbulent fluxes in plant The forest begins to warm and some of this
canopies (Baldocchi 1992; Baldocchi and Meyers energy is returned to the atmosphere as sensi-
1998; Pyles et al. 2000). ble and latent heat. Stomata open, allowing for
net CO2 uptake during photosynthesis. Fluxes
are strongest in early to middle afternoon and
17.8 Environmental Controls of decrease late in the afternoon when solar radia-
tion diminishes.
Canopy Fluxes The concentration of CO2 in the near-surface
atmosphere can have a strong diurnal cycle in
Surface fluxes vary over the course of a day in response to CO2 uptake during the day and loss
response to the diurnal cycle. As more radiation during the night, as seen in measurements taken
is received from the Sun, the land warms and at the boreal aspen forest (Figure 17.15). In the
more energy is dissipated as sensible and latent dormant season, when trees are leafless, there is
heat. In addition, stomata open to allow CO2 little diurnal variation in CO2 concentration; nor
uptake, but in doing so plants lose water during is there a strong vertical gradient. In the grow-
transpiration. Figure 12.7 illustrates these cycles ing season, however, there is a strong diurnal
for boreal aspen and jack pine forests on a typ- cycle at heights of 0.8 and 2.3 m. Nighttime con-
ical summer day. In early morning, the forest centrations range from 400 to 500 ppm while

Downloaded from https://www.cambridge.org/core. Danish National Library of Science and Medicine, on 25 Apr 2022 at 11:50:28, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781107339200.018
282 Plant Canopies

30 Fig. 17.16 Net ecosystem
exchange of carbon (NEE)
20 in relation to direct (clear
Cloudy days days) and diffuse (cloudy days)
Clear days photosynthetically active radiation
10
NEE (µmol CO2 m–2 s–1)

(PPFD). Data are for a boreal


aspen forest. Negative fluxes
0 indicate carbon uptake. Adapted
from Law et al. (2002).
–10

–20

–30

–40
0 500 1000 1500 2000
PPFD (µmol m–2 s–1)

daytime concentrations are about 350 ppm. The radiation was diffuse, than for clear sky, when
high nighttime concentrations arise from respi- the radiation was primarily direct beam. Similar
ration losses and the occurrence of a strong tem- enhancement of CO2 uptake when radiation is
perature inversion that suppresses the upward predominantly diffuse compared with when it
turbulent transport of CO2. The magnitude of is predominantly direct beam has been found
this diurnal cycle diminishes with height so that in other forests (Hollinger et al. 1994; Goulden
at a height of 9.5 m and above, nighttime and et  al. 1997; Gu et  al. 2002, 2003; Knohl and
daytime CO2 concentrations are similar. Baldocchi 2008).
Canopy fluxes show similar functional rela- Similar to leaf fluxes, GPP increases as evapo-
tionships with environmental controls as do transpiration (ET) increases (Figure 17.17). At the
leaf fluxes. For example, the net uptake of car- canopy-scale, GPP / ET is a measure of water-use
bon by a forest canopy increases as photosyn- efficiency; this quantity scaled by vapor pressure
thetically active radiation increases, saturating deficit (GPP ∗ VPD / ET ) is the canopy equivalent
at high light levels (Figure 17.16). This relation- of intrinsic water-use efficiency. Among vari-
ship can be described by: ous forests, grasslands, and croplands, GPP / ET
ranges from 1 to 6 g C kg–1 H2O; GPP ∗ VPD / ET
A max ΕF ↓ ranges from 5 to 43 g C hPa kg–1 H2O (Beer
NEE = − + R E (17.25)
ΕF ↓ + A max et al. 2009).
In deciduous forests, the emergence of leaves
The first term on the right-hand side of this in spring alters the partitioning of net radiation
equation is gross primary production (GPP) and into sensible and latent heat fluxes. Table  17.3
is the canopy-scale equivalent of Eq. (17.8); the shows the effects of leaf emergence in a tem-
second term is ecosystem respiration (Lasslop perate deciduous forest. The majority of net
et al. 2010). At the canopy scale, carbon uptake radiation was dissipated as sensible heat when
can vary depending on the amount of direct the canopy was leafless. After leaf emergence,
beam and diffuse radiation. Diffuse radiation latent heat was the dominant flux. The Bowen
penetrates more deeply into a canopy than does ratio above the canopy decreased from 2.1 dur-
direct beam radiation. In the particular for- ing the dormant season to 0.5 during the grow-
est illustrated in Figure  17.16, net CO2 uptake ing season due to a large increase in the fraction
was greater for cloudy skies, when more of the of net radiation partitioned into latent heat

Downloaded from https://www.cambridge.org/core. Danish National Library of Science and Medicine, on 25 Apr 2022 at 11:50:28, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781107339200.018
17.8 Environmental Controls of Canopy Fluxes 283

500 Fig. 17.17  Monthly gross


primary production (GPP)
in relation to monthly
400
evapotranspiration (ET) for
broadleaf deciduous forests.
300 Adapted from Law et al. (2002).
GPP (g CO2 m–2)

200

100

–100
0 20 40 60 80 100 120

ET (kg H2O m ) –2

Table 17.3  Dormant season and growing season energy fluxes above the canopy, at the forest floor, and from
vegetation (canopy – floor)

Canopy Floor Vegetation Floor (%)

Dormant season
Rn (MJ m–2) 720 315 405 44
H (MJ m–2) 419 193 226 46
λE (MJ m–2) 200 112 88 56
H / λE 2.1 1.7 2.6 –
Growing season
Rn (MJ m–2) 2080 287 1793 14
H (MJ m–2) 617 38 579 6
λE (MJ m–2) 1140 95 1045 8
H / λE 0.5 0.4 0.6 –
Note: Rn , net radiation; λE , latent heat flux; H , sensible heat flux. Fluxes are summed over the dormant season (days 1–115,
305–365) and growing season (days 116–304). The forest floor contribution to canopy flux is also shown as a percentage.
Source: From Wilson et al. (2000).

flux. The emergence of leaves also altered fluxes than 10  percent to canopy sensible and latent
at the forest floor. In the dormant season, net heat fluxes.
radiation at the forest floor was 44  percent of Leaf area index is a critical determinant of
that above the canopy. Sensible and latent heat the proportion of evapotranspiration that is
fluxes at the forest floor contributed 46 percent transpiration, and this ratio can be 0.8–0.9 or
and 56 percent, respectively, of the total canopy higher in closed canopies (Wang et al. 2014). The
fluxes. After leaf emergence, net radiation at the radiation profile influences the partitioning of
forest floor was only 14  percent of that above total evapotranspiration into transpiration from
the canopy, and the forest floor contributed less foliage and evaporation from the forest floor. In

Downloaded from https://www.cambridge.org/core. Danish National Library of Science and Medicine, on 25 Apr 2022 at 11:50:28, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781107339200.018
284 Plant Canopies

(a) 10 (b) 10

Forest (needleleaf) R2 = 0.76


Forest (broadleaf)
Shrub
1 Grass 1
Crop
Roughness length (m)

Roughness length (m)


0.1 0.1

Forest (needleleaf)
Forest (broadleaf)
Shrub
0.01 0.01 Grass
Crop
Desert
R2 = 0.74
0.001 0.001
0.01 0.1 1 10 0.00 0.05 0.10 0.15 0.20 0.25 0.30
Height /LAI Albedo

Fig. 17.18  Relationships among roughness length, canopy structure, and albedo for forest, shrub, grass, crop, and semidesert
vegetation. Roughness length is shown in relation to (a) the ratio of canopy height to leaf area index and (b) albedo.
Measurements were made at peak leaf area and albedo is at mid-day. Data from Cho et al. (2012).

general, the relative contribution of the forest Canopy structure influences canopy aerody-
floor and understory to total evapotranspiration namics. Simplified relationships specify rough-
is relatively little (~10%) in dense forests with ness length in relation to canopy height (h) by
high leaf area index and greater (up to 50%) in z 0 m = 0.1h and similarly displacement height by
open canopy forests, where the canopy absorbs d = 0.7 h (Chapter  13). Other parameterizations
less radiation (Baldocchi and Ryu 2011). This can additionally include a dependence on leaf area
be seen in Figure  17.9, where the surface con- index and other factors (Raupach 1994; Sellers
ductance is greater than canopy conductance et al. 1996a). In an analysis of forest, grassland,
at low leaf area index and approaches canopy cropland, and shrubland vegetation, Cho et  al.
conductance at high leaf area index and is seen (2012) found that roughness length correlates
also in the effects of leaf emergence in decidu- with the ratio of canopy height to leaf area
ous forests (Table 17.3). index (h / L) (Figure  17.18a). Forests have large
The relationship between forest floor latent roughness length and high h / L. The shorter
heat flux and available energy at the forest statured grasses, crops, and shrubs have smaller
floor observed for a variety of forests illustrates roughness length and lower h / L. The shorter
the strong control exerted by available energy vegetation also has higher surface albedo so
(Baldocchi et al. 2000). When the forest floor is that albedo and roughness length are negatively
dry, its latent heat flux increases linearly with correlated (Figure  17.18b). This latter relation-
greater available energy and accounts for about ship illustrates a tradeoff between albedo and
25  percent of available energy. At high avail- roughness length, with different consequences
able energy beyond about 100 W m–2, however, for net radiation (albedo) and turbulent fluxes
forest floor evaporation is not able to meet the (roughness length).
potential demand and latent heat flux reaches a
threshold of about 35 W m–2.

17.9  Review Questions

1. Use Eq. (17.1) to calculate the amount of the canopy) reaching the forest floor for a leaf area
solar radiation (as a percentage of that at the top of index of 4 m2 m–2 and spherical, horizontal, and

Downloaded from https://www.cambridge.org/core. Danish National Library of Science and Medicine, on 25 Apr 2022 at 11:50:28, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781107339200.018
17.10 References 285

vertical foliage. Calculations are for a declination for Ta = 25°C), and g ah = 2 mol m–2 s–1. Relative humid-
angle δ = 23°27′, latitude φ = 40° N, and time of day is ity is 75 percent. Use c p = 29.2 J mol–1 K–1 and γ  = 66.5
1200 hours and 1500 hours. Zenith angle is calculated Pa K–1.
from Eq. (4.1). How does zenith angle interact with 7. Discuss differences and/or similarities among
leaf orientation to affect radiative transfer? leaf conductance, surface conductance, and canopy
2. What is the maximum sunlit leaf area in a can- conductance in a deciduous forest when the can-
opy with spherically oriented foliage? opy is leafless. How does leaf emergence affect these
3. The maximum absorption of solar radiation conductances?
in the visible waveband is about 95 percent. Use the 8. Temperature is 30.8°C at the forest floor,
production efficiency model to calculate daily canopy 31.2°C in the understory, 30.1°C in the overstory,
photosynthesis if incoming solar radiation averaged and 28.3°C above the canopy. Sensible heat flux is 60
over the day is 160 W m–2. Assume solar radia- W m–2 at the forest floor and 300 W m–2 above the
tion is 50  percent visible waveband and 50  percent canopy. Are these indicative of gradient–diffusion or
near-infrared and that environmental factors do not counter-gradient fluxes?
limit photosynthesis. 9. A surface emits longwave radiation of 500 W
4. Contrast the production efficiency model with m–2. What is the radiative temperature for an emissiv-
the GPP model given by Eq. (17.25). What is a key dif- ity of 1.0 and 0.95?
ference between these two models? Do they differ in 10. Measured sensible heat flux is H = 200 W
their timescale? m–2. One student evaluates surface temperature with
5. Use Eq. (17.9) to calculate canopy photosyn- z 0 m = 0.5 m. Another student uses z 0 m = 0.2 m. What
thesis for spherical, horizontal, and vertical foliage. is the difference in estimated surface temperature?
L  = 6 m2 m–2, A max  = 20 μmol m–2 s–1, Ε  = 0.06, and F0= Use c p = 29.2 J mol–1 K–1, ρm = 42.3 mol m–3, and u =
1000 μmol m–2 s–1. Zenith angle is Z = 30°. Repeat the *
0.2 m s–1.
calculation for Z = 60°. How does leaf orientation inter- 11.  Calculate the change in air temperature in a
act with zenith angle to affect canopy photosynthesis? 20 m tall canopy with H v = 85 W m–2, H g  = 8 W m–2,
6. Use the Penman–Monteith equation to calcu- and H = 83 W m–2. How much heat is stored in the
late canopy conductance for the following conditions canopy? Use ρm = 42.3 mol m–3 and c p = 29.2 J mol–1 K–1.
with latent heat flux equal to 300 W m–2:  R n − G = What is the change in temperature if H = 93 W m–2?
400 W m–2, e* (Ta )= 3167 Pa and s = 189 Pa K–1 (values

17.10 References

Baldocchi, D. D. (1989). Turbulent transfer in a decidu- Baldocchi, D. D., and Meyers, T. P. (1998). On using
ous forest. Tree Physiology, 5, 357–377. eco-physiological, micrometeorological and biogeo-
Baldocchi, D. D. (1992). A Lagrangian random-walk chemical theory to evaluate carbon dioxide, water
model for simulating water vapor, CO2 and sensi- vapor and trace gas fluxes over vegetation:  a per-
ble heat flux densities and scalar profiles over and spective. Agricultural and Forest Meteorology, 90, 1–25.
within a soybean canopy. Boundary-Layer Meteorology, Baldocchi, D. D., and Ryu, Y. (2011). A synthesis of for-
61, 113–144. est evaporation fluxes – from days to years – as mea-
Baldocchi, D. D., and Meyers, T. P. (1988a). A spec- sured with eddy covariance. In Forest Hydrology and
tral and lag-correlation analysis of turbulence in a Biogeochemistry:  Synthesis of Past Research and Future
deciduous forest canopy. Boundary-Layer Meteorology, Directions, ed. D. F. Levia, D. Carlyle-Moses, and T.
45, 31–58. Tanaka. Dordrecht: Springer, pp. 101–116.
Baldocchi, D. D., and Meyers, T. P. (1988b). Turbulence Baldocchi, D. D., and Xu, L. (2007). What limits evap-
structure in a deciduous forest. Boundary-Layer oration from Mediterranean oak woodlands  – the
Meteorology, 43, 345–364. supply of moisture in the soil, physiological con-
Baldocchi, D. D., and Meyers, T. P. (1989). The effects trol by plants or the demand by the atmosphere?
of extreme turbulent events on the estimation of Advances in Water Resources, 30, 2113–2122.
aerodynamic variables in a deciduous forest can- Baldocchi, D. D., Vogel, C. A., and Hall, B. (1997).
opy. Agricultural and Forest Meteorology, 48, 117–134. Seasonal variation of energy and water vapor

Downloaded from https://www.cambridge.org/core. Danish National Library of Science and Medicine, on 25 Apr 2022 at 11:50:28, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781107339200.018
286 Plant Canopies

exchange rates above and below a boreal jack pine The Forest–Atmosphere Interaction, ed. B. A. Hutchinson
forest canopy. Journal of Geophysical Research, 102D, and B. B. Hicks. Dordrecht: Reidel, pp. 421–442.
28939–28951. de Pury, D. G. G., and Farquhar, G. D. (1997). Simple
Baldocchi, D. D., Law, B. E., and Anthoni, P. M. (2000). scaling of photosynthesis from leaves to canopies
On measuring and modeling energy fluxes above without the errors of big-leaf models. Plant, Cell and
the floor of a homogeneous and heterogeneous Environment, 20, 537–557.
conifer forest. Agricultural and Forest Meteorology, 102, Dickinson, R. E., Henderson-Sellers, A., Kennedy,
187–206. P. J., and Wilson, M.F. (1986). Biosphere–Atmosphere
Beer, C., Ciais, P., Reichstein, M., et al. (2009). Temporal Transfer Scheme (BATS) for the NCAR Community Climate
and among-site variability of inherent water use effi- Model, Technical Note NCAR/TN-275+STR. Boulder,
ciency at the ecosystem level. Global Biogeochemical Colorado:  National Center for Atmospheric
Cycles, 23, GB2018, doi:10.1029/2008GB003233. Research.
Beljaars, A. C. M., and Holtslag, A. A. M. (1991). Flux Dickinson, R. E., Henderson-Sellers, A., and Kennedy,
parameterization over land surfaces for atmo- P. J. (1993). Biosphere–Atmosphere Transfer Scheme
spheric models. Journal of Applied Meteorology, 30, (BATS) Version 1e as Coupled to the NCAR Community
327–341. Climate Model, Technical Note NCAR/TN-387+STR.
Bonan, G. B., Lawrence, P. J., Oleson, K. W., et al. (2011). Boulder, Colorado: National Center for Atmospheric
Improving canopy processes in the Community Research.
Land Model version 4 (CLM4) using global flux Dorman, J. L., and Sellers, P. J. (1989). A global clima-
fields empirically inferred from FLUXNET data. tology of albedo, roughness length and stomatal
Journal of Geophysical Research, 116, G02014, resistance for atmospheric general circulation mod-
doi:10.1029/2010JG001593. els as represented by the simple biosphere model
Campbell, G. S., and Norman, J. M. (1998). An (SiB). Journal of Applied Meteorology, 28, 833–855.
Introduction to Environmental Biophysics, 2nd ed. Farquhar, G. D., von Caemmerer, S., and Berry, J. A.
New York: Springer-Verlag. (1980). A biochemical model of photosynthetic
Carswell, F. E., Meir, P., Wandelli, E. V., et al. (2000). CO2 assimilation in leaves of C3 species. Planta,
Photosynthetic capacity in a central Amazonian 149, 78–90.
rain forest. Tree Physiology, 20, 179–186. Finnigan, J. (2000). Turbulence in plant canopies.
Cho, J., Miyazaki, S., Yeh, P. J.-F., et al. (2012). Testing Annual Review of Fluid Mechanics, 32, 519–571.
the hypothesis on the relationship between aero- Garratt, J. R. (1978). Transfer characteristics for a het-
dynamic roughness length and albedo using vege- erogeneous surface of large aerodynamic rough-
tation structure parameters. International Journal of ness. Quarterly Journal of the Royal Meteorological
Biometeorology, 56, 411–418. Society, 104, 491–502.
Choudhury, B. J., and Monteith, J. L. (1988). A four-layer Garratt, J. R., and Hicks, B. B. (1973). Momentum,
model for the heat budget of homogeneous land heat and water vapour transfer to and from natu-
surfaces. Quarterly Journal of the Royal Meteorological ral and artificial surfaces. Quarterly Journal of the Royal
Society, 114, 373–398. Meteorological Society, 99, 680–687.
Dai, Y., Dickinson, R. E., and Wang, Y.-P. (2004). A Goudriaan, J. (1977). Crop Micrometeorology: A Simulation
two-big-leaf model for canopy temperature, pho- Study. Wageningen:  Center for Agricultural
tosynthesis, and stomatal conductance. Journal of Publication and Documentation.
Climate, 17, 2281–2299. Goudriaan, J., and van Laar, H. H. (1994). Modelling
Dang, Q. L., Margolis, H. A., Sy, M., et al. (1997). Profiles Potential Crop Growth Processes: Textbook with Exercises.
of photosynthetically active radiation, nitrogen and Dordrecht: Kluwer.
photosynthetic capacity in the boreal forest: impli- Goulden, M. L., Daube, B. C., Fan, S.-M., et al. (1997).
cations for scaling from leaf to canopy. Journal of Physiological responses of a black spruce forest
Geophysical Research, 102D, 28845–28859. to weather. Journal of Geophysical Research, 102D,
Deardorff, J. W. (1978). Efficient prediction of ground 28987–28996.
surface temperature and moisture, with inclusion Grace, J. (1988). The functional significance of short
of a layer of vegetation. Journal of Geophysical Research, stature in montane vegetation. In Plant Form and
83C, 1889–1903. Vegetation Structure, ed. M. J. A. Werger, P. J. M. van
Denmead, O. T., and Bradley, E. F. (1985). der Aart, H. J. During, and J. T.  A. Verhoeven. The
Flux–gradient relationships in a forest canopy. In Hague: SPB Academic Publishing, pp. 201–209.

Downloaded from https://www.cambridge.org/core. Danish National Library of Science and Medicine, on 25 Apr 2022 at 11:50:28, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781107339200.018
17.10 References 287

Grace, J., Allen, S. J., and Wilson, C. (1989). Climate within and above a deciduous forest. Agricultural and
and the meristem temperatures of plant communi- Forest Meteorology, 53, 207–222.
ties near the tree-line. Oecologia, 79, 198–204. Meyers, T., and Paw U, K. T. (1986). Testing of a
Gu, L., Baldocchi, D., Verma, S. B., et  al. (2002). higher-order closure model for modeling airflow
Advantages of diffuse radiation for terrestrial eco- within and above plant canopies. Boundary-Layer
system productivity. Journal of Geophysical Research, Meteorology, 37, 297–311.
107, 4050, 10.1029/2001JD001242. Meyers, T., and Paw U, K. T. (1987). Modelling the
Gu, L., Baldocchi, D. D., Wofsy, S. C., et  al. (2003). plant canopy micrometeorology with higher-order
Response of a deciduous forest to the Mount closure principles. Agricultural and Forest Meteorology,
Pinatubo eruption:  enhanced photosynthesis. 41, 143–163.
Science, 299, 2035–2038. Monteith, J. L. (1965). Evaporation and environment.
Hollinger, D. Y. (1996). Optimality and nitrogen allo- In The State and Movement of Water in Living Organisms
cation in a tree canopy. Tree Physiology, 16, 627–634. (19th Symposia of the Society for Experimental Biology),
Hollinger, D. Y., Kelliher, F. M., Byers, J. N., et al. (1994). ed. G. E. Fogg. New  York:  Academic Press, pp.
Carbon dioxide exchange between an undisturbed 205–234.
old-growth temperate forest and the atmosphere. Monteith, J. L. (1977). Climate and the efficiency of
Ecology, 75, 134–150. crop production in Britain. Philosophical Transactions
Jarvis, P. G., and McNaughton, K. G. (1986). Stomatal of the Royal Society B, 281, 277–294.
control of transpiration:  scaling up from leaf to Myneni, R. B., Keeling, C. D., Tucker, C. J., Asrar, G.,
region. Advances in Ecological Research, 15, 1–49. and Nemani, R. R. (1997). Increased plant growth
Kelliher, F. M., Leuning, R., Raupach, M. R., and in the northern high latitudes from 1981 to 1991.
Schulze, E.-D. (1995). Maximum conductances Nature, 386, 698–702.
for evaporation from global vegetation types. Niinemets, Ü. (2007). Photosynthesis and resource
Agricultural and Forest Meteorology, 73, 1–16. distribution through plant canopies. Plant, Cell and
Knohl, A., and Baldocchi, D. D. (2008). Effects of dif- Environment, 30, 1052–1071.
fuse radiation on canopy gas exchange processes Norman, J. M. (1979). Modeling the complete crop
in a forest ecosystem. Journal of Geophysical Research, canopy. In Modification of the Aerial Environment of
113, G02023, doi:10.1029/2007JG000663. Plants, ed. B. J. Barfield and J. F. Gerber. St. Joseph,
Körner, C. (1994). Leaf diffusive conductances in the Michigan:  American Society of Agricultural
major vegetation types of the globe. In Ecophysiology Engineers, pp. 249–277.
of Photosynthesis, ed. E.-D. Schulze and M. M. Caldwell. Norman, J. M. (1993). Scaling processes between
New York: Springer-Verlag, pp. 463–490. leaf and canopy levels. In Scaling Physiological
Lasslop, G., Reichstein, M., Papale, D., et  al. (2010). Processes: Leaf to Globe, ed. J. R. Ehleringer and C. B.
Separation of net ecosystem exchange into assim- Field. New York: Academic Press, pp. 41–76.
ilation and respiration using a light response curve Prince, S. D., and Goward, S. N. (1995). Global primary
approach:  Critical issues and global evaluation. production:  a remote sensing approach. Journal of
Global Change Biology, 16, 187–208. Biogeography, 22, 815–835.
Law, B. E., Falge, E., Gu, L., et al. (2002). Environmental Pyles, R. D., Weare, B. C., and Paw U, K. T. (2000).
controls over carbon dioxide and water vapor The UCD advanced canopy–atmosphere–soil algo-
exchange of terrestrial vegetation. Agricultural and rithm: Comparisons with observations from differ-
Forest Meteorology, 113, 97–120. ent climate and vegetation regimes. Quarterly Journal
Lloyd, J., Patiño, S., Paiva, R. Q., et  al. (2010). of the Royal Meteorological Society, 126, 2951–2980.
Optimisation of photosynthetic carbon gain and Rauner, J. L. (1976). Deciduous forests. In Vegetation and
within-canopy gradients of associated foliar traits the Atmosphere: vol. 2. Case Studies, ed. J. L. Monteith.
for Amazon forest trees. Biogeosciences, 7, 1833–1859. New York: Academic Press, pp. 241–264.
Meir, P., Kruijt, B., Broadmeadow, M., et  al. (2002). Raupach, M. R. (1994). Simplified expressions for veg-
Acclimation of photosynthetic capacity to irradi- etation roughness length and zero-plane displace-
ance in tree canopies in relation to leaf nitrogen ment as functions of canopy height and area index.
concentration and leaf mass per unit area. Plant, Cell Boundary-Layer Meteorology, 71, 211–216.
and Environment, 25, 343–357. Raupach, M. R., and Finnigan, J. J. (1988). Single-layer
Meyers, T. P., and Baldocchi, D. D. (1991). The budgets models of evaporation from plant canopies are
of turbulent kinetic energy and Reynolds stress incorrect but useful, whereas multilayer models are

Downloaded from https://www.cambridge.org/core. Danish National Library of Science and Medicine, on 25 Apr 2022 at 11:50:28, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781107339200.018
288 Plant Canopies

correct but useless: Discuss. Australian Journal of Plant Tucker, C. J., Fung, I. Y., Keeling, C. D., and Gammon,
Physiology, 15, 705–716. R. H. (1986). Relationship between atmospheric CO2
Rogers, A. (2014). The use and misuse of Vc,max in Earth variations and a satellite-derived vegetation index.
System Models. Photosynthesis Research, 119, 15–29. Nature, 319, 195–199.
Running, S. W., Thornton, P. E., Nemani, R., and Tucker, C. J., Dregne, H. E., and Newcomb, W. W.
Glassy, J. M. (2000). Global terrestrial gross and net (1991). Expansion and contraction of the Sahara
primary productivity from the Earth Observing Desert from 1980 to 1990. Science, 253, 299–301.
System. In Methods in Ecosystem Science, ed. O. E. Sala. Vidale, P. L., and Stöckli, R. (2005). Prognostic
New York: Springer-Verlag, pp. 44–57. canopy air space solutions for land surface
Running, S. W., Nemani, R. R., Heinsch, F. A., et al. (2004). exchanges. Theoretical and Applied Climatology, 80,
A continuous satellite-derived measure of global ter- 245–257.
restrial primary production. BioScience, 54, 547–560. Wang, L., Good, S. P., and Caylor, K. K. (2014). Global
Schulze, E.-D., Kelliher, F. M., Körner, C., Lloyd, J., and synthesis of vegetation control on evapotranspi-
Leuning, R. (1994). Relationships among maximum ration partitioning. Geophysical Research Letters, 41,
stomatal conductance, ecosystem surface conduc- 6753–6757, doi:10.1002/2014GL061439.
tance, carbon assimilation rate, and plant nitrogen Wang, Y.-P., and Leuning, R. (1998). A two-leaf model
nutrition: A global ecology scaling exercise. Annual for canopy conductance, photosynthesis and parti-
Review of Ecology and Systematics, 25, 629–660. tioning of available energy, I: Model description and
Sellers, P. J. (1985). Canopy reflectance, photosynthe- comparison with a multi-layered model. Agricultural
sis and transpiration. International Journal of Remote and Forest Meteorology, 91, 89–111.
Sensing, 6, 1335–1372. Wilson, C., Grace, J., Allen, S., and Slack, F. (1987).
Sellers, P. J., Mintz, Y., Sud, Y. C., and Dalcher, A. (1986). Temperature and stature:  a study of tempera-
A simple biosphere model (SiB) for use within gen- tures in montane vegetation. Functional Ecology, 1,
eral circulation models. Journal of the Atmospheric 405–414.
Sciences, 43, 505–531. Wilson, J. D. (1988). A second-order closure model for
Sellers, P. J., Berry, J. A., Collatz, G. J., Field, C. B., and flow through vegetation. Boundary-Layer Meteorology,
Hall, F. G. (1992). Canopy reflectance, photosyn- 42, 371–392.
thesis, and transpiration, III:  A  reanalysis using Wilson, K. B., Hanson, P. J., and Baldocchi, D. D.
improved leaf models and a new canopy integration (2000). Factors controlling evaporation and energy
scheme. Remote Sensing of Environment, 42, 187–216. partitioning beneath a deciduous forest over an
Sellers, P. J., Los, S. O., Tucker, C. J., et  al. (1996a). A annual cycle. Agricultural and Forest Meteorology, 102,
revised land surface parameterization (SiB2) for 83–103.
atmospheric GCMs, Part II: The generation of global Wilson, N. R., and Shaw, R. H. (1977). A higher order
fields of terrestrial biophysical parameters from sat- closure model for canopy flow. Journal of Applied
ellite data. Journal of Climate, 9, 706–737. Meteorology, 16, 1197–1205.
Sellers, P. J., Randall, D. A., Collatz, G. J., et al. (1996b). Yang, P. C., Black, T. A., Neumann, H. H., Novak, M.
A revised land surface parameterization (SiB2) for D., and Blanken, P. D. (1999). Spatial and temporal
atmospheric GCMs, Part I:  Model formulation. variability of CO2 concentration and flux in a boreal
Journal of Climate, 9, 676–705. aspen forest. Journal of Geophysical Research, 104D,
Shuttleworth, W. J., and Wallace, J. S. (1985). 27653–27661.
Evaporation from sparse crops – an energy combina- Yuan, W., Liu, S., Zhou, G., et  al. (2007). Deriving a
tion theory. Quarterly Journal of the Royal Meteorological light use efficiency model from eddy covariance
Society, 111, 839–855. flux data for predicting daily gross primary produc-
Thom, A. S. (1972). Momentum, mass and heat tion across biomes. Agricultural and Forest Meteorology,
exchange of vegetation. Quarterly Journal of the Royal 143, 189–207.
Meteorological Society, 98, 124–134. Zhao, M., Heinsch, F. A., Nemani, R. R., and Running,
Tucker, C. J., Townshend, J. R. G., and Goff, T. E. (1985). S. W. (2005). Improvements of the MODIS terrestrial
African land-cover classification using satellite data. gross and net primary production global data set.
Science, 227, 369–375. Remote Sensing of Environment, 95, 164–176.

Downloaded from https://www.cambridge.org/core. Danish National Library of Science and Medicine, on 25 Apr 2022 at 11:50:28, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781107339200.018

You might also like