ArticlePDF Available

Facultative and constitutive pigment effects on the Photochemical Reflectance Index (PRI) in sun and shade conifer needles

Authors:

Abstract and Figures

Leaf pigment content and spectral reflectance were examined in four conifer species from the Pacific Northwest and Canadian boreal forest. Our goal was to evaluate the causes of within- and between-stand variation in the Photochemical Reflectance Index (PRI), an indicator of xanthophyll cycle activity and carotenoid pigment content that often scales with photosynthetic light-use efficiency. Both the dark-state PRI values and the change in PRI upon dark-light transition (ΔPRI) were measured in situ in leaves from different canopy positions (top vs. bottom) having contrasting light histories (sun vs. shade). PRI varied with species, canopy position, and with the pool sizes of several photoprotective carotenoid pigments (relative to chlorophyll). Upper-canopy leaves had a greater Δ PRI than their shaded counterparts lower in the canopy, reflecting a higher investment of the photoprotective xanthophyll cycle pigments for sun-exposed top-canopy leaves. These results indicate that the relative concentration of different pigment groups and associated PRI responses varied with canopy position and light history over more than one time scale, and included rapidly changing (facultative) and slowly changing (constitutive) components. Most of the PRI variability among the forest trees sampled was due to constitutive pigment pool size variation associated with species and canopy position. We conclude that both facultative and constitutive pigment components should be considered when applying PRI to photosynthetic studies of forest stands with remote sensing. Leaf-level measurements of PRI and ΔPRI provide non-destructive probes of both facultative and constitutive pigment changes within plant canopies that could help interpret variation in PRI signal viewed from remote sensing platforms.
Content may be subject to copyright.
© 2012 Science From Israel / LPPLtd., Jerusalem
Israel Journal of Plant Sciences Vol. 60 2012 pp. 85–95
DOI: 10.1560/IJPS.60.1-2.85
*Author to whom correspondence should be addressed.
E-mail: jgamon@gmail.com
Facultative and constitutive pigment effects on the Photochemical Reectance Index
(PRI) in sun and shade conifer needles
John A. GAmona,* And Joseph A. Berryb
aDepartments of Earth and Atmospheric Sciences & Biological Sciences, University of Alberta, Edmonton,
Alberta T6G 2E3, Canada
bDepartment of Global Ecology, Carnegie Institution for Science, Stanford, California 94305, USA
(Received 30 January 2012; accepted in revised form 12 February 2012)
Honoring Anatoly Gitelson on the occasion of his 70th birthday
ABSTRACT
Leaf pigment content and spectral reectance were examined in four conifer species
from the Pacic Northwest and Canadian boreal forest. Our goal was to evaluate
the causes of within- and between-stand variation in the Photochemical Reectance
Index (PRI), an indicator of xanthophyll cycle activity and carotenoid pigment con-
tent that often scales with photosynthetic light-use efciency. Both the dark-state
PRI values and the change in PRI upon dark–light transition (ΔPRI) were measured
in situ in leaves from different canopy positions (top vs. bottom) having contrasting
light histories (sun vs. shade). PRI varied with species, canopy position, and with the
pool sizes of several photoprotective carotenoid pigments (relative to chlorophyll).
Upper-canopy leaves had a greater Δ PRI than their shaded counterparts lower in
the canopy, reecting a higher investment of the photoprotective xanthophyll cycle
pigments for sun-exposed top-canopy leaves. These results indicate that the rela-
tive concentration of different pigment groups and associated PRI responses varied
with canopy position and light history over more than one time scale, and included
rapidly changing (facultative) and slowly changing (constitutive) components. Most
of the PRI variability among the forest trees sampled was due to constitutive pig-
ment pool size variation associated with species and canopy position. We conclude
that both facultative and constitutive pigment components should be considered
when applying PRI to photosynthetic studies of forest stands with remote sensing.
Leaf-level measurements of PRI and ΔPRI provide non-destructive probes of both
facultative and constitutive pigment changes within plant canopies that could help
interpret variation in PRI signal viewed from remote sensing platforms.
Keywords: Photochemical Reectance Index (PRI), leaf pigments, irradiance, coni-
fers, carotenoids, xanthophyll cycle
INTRODUCTION
Plant leaves balance their requirements for light harvest-
ing and photoprotection many ways, including chloro-
phlast movement (Brugnoli and Björkman, 1992; Park
et al., 1996; Zygielbaum et al., 2012), leaf movement
(Gamon and Pearcy, 1989; Jiang et al., 2006), pigment
pool size changes (Thayer and Bjorkman, 1990; García-
Plazaola et al., 1997; Demmig-Adams, 1998; Stylinski
et al., 2002), and diurnally changing energy distribution
through xanthophyll pigment conversion and non-
photochemical dissipation of absorbed photons (Dem-
mig-Adams and Adams, 1992). Slow or irreversible
adjustments are often termed constitutive properties,
whereas rapid, reversible adjustments can be considered
Israel Journal of Plant Sciences 60 2012
86
facultative. Examples of facultative responses include
the changing epoxidation state of the xanthophyll
cycle pigments in response to diurnally changing irradi-
ance (Demmig-Adams and Adams, 1992). Examples
of constitutive adjustments include enhanced levels
(pool sizes) of carotenoid pigments with increased sun
exposure (Demmig-Adams, 1998; Niinemets et al.,
2003). Similarly, pigment pool sizes change gradually
during leaf development (Gamon and Surfus, 1999),
seasonal change (García-Plazaola et al., 1997; Stylin-
ski et al., 2002), or leaf age (Gitelson and Merzlyak,
1994; Merzlyak and Gitelson, 1995). In this study, we
dene “facultative effects” as those that change on a
daily or shorter time scale, and “constitutive effects” as
those that change over much longer time scales (e.g.,
due to changing pigment pool sizes in response to sun
exposure, leaf age, or chronic stress). These facultative
and constitutive pigment effects both inuence leaf
reectance properties and can be monitored with the
appropriate “reectance index.”
Many leaf pigment indices have been developed to
measure chlorophyll, carotenoid, or anthocyanin levels
using spectral reectance (Gamon and Surfus, 1999;
Merzlyak et al., 1999, 2003; Sims and Gamon, 2002;
Gitelson et al., 2006, 2009; Steele et al., 2009; Ustin
et al., 2009). When hyperspectral detectors comprised
of many narrow wavebands are used, spectral reec-
tance can resolve narrow-band features diagnostic of
individual pigments, pigment groups, or relative levels
of major pigment groups (Ustin et al., 2009). With the
appropriate foreoptics, these instruments can be used for
assessing leaf pigments in the eld. A particular advan-
tage of this approach is that it provides a rapid, non-de-
structive in situ sampling method that can preclude the
need for extensive and costly laboratory analyses.
The Photochemical Reectance Index (PRI) was
originally developed as an optical measure of leaf
xanthophyll cycle pigment activity that detects the in-
terconversion of xanthophyll cycle pigments in leaves
and plant stands over diurnal time spans (Gamon et al.,
1992, 1993; Peñuelas et al., 1995). Because xanthophyll
pigment conversion is closely linked to non-photo-
chemical quenching and PSII photochemical efciency
(Demmig-Adams and Adams, 1992), PRI can be related
to leaf- and stand-level photosynthetic activity and of-
fers promise as a “scaleable” remote sensing measure of
photosynthetic rate or light-use efciency (LUE) over
diurnal time scales (Gamon et al., 1992 and 2001).
In recent years, many tests of PRI as a remote in-
dicator of stand- or ecosystem-level photosynthetic
activity have been conducted using aircraft or satellite
measurements, and several have found signicant cor-
relations between PRI and whole-ecosystem photosyn-
thetic activity or LUE measured using eddy covariance
(Nichol et al., 2000; Rahman et al., 2001, 2004; Drolet
et al., 2005; Garbulsky et al., 2008; Hilker et al., 2009;
Goerner et al., 2011). These encouraging results have
led to much recent speculation that satellite-based mea-
sures employing PRI (or a similar index) could be used
to improve remote sensing of “photosynthesis from
space” (Grace et al., 2007; Coops et al., 2010; Hall et
al., 2011).
Remote sensing studies of PRI as a measure of stand
or ecosystem photosynthetic activity generally have not
considered the potential for constitutive changes in leaf
pigment content inuencing the observed PRI. Since
most eld and remote sensing studies do not actually
measure pigment levels, they often assume that varia-
tion in PRI primarily reects short-term xanthophyll
cycle pigment activity, and ignore the potential contri-
bution of constitutive pool size differences. However,
many long-term eld studies have now shown that PRI
can also be strongly affected by changing pigment pool
sizes, particularly over seasonal time spans or across
species (Stylinski et al., 2002; Filella et al., 2004,
2009; Sims et al., 2006), In some ecosystems, these
constitutive effects can be the primary source of PRI
variation, and are linked to prevailing environmental
conditions (Sims et al., 2006), which can involve irradi-
ance (Niinemets et al., 2003), temperature (Sims et al.,
2006), nutrients (Gamon et al., 1997), and water status
(Sims et al., 2006). Currently, the effect of constitutive
pigment changes on the PRI–LUE relationship has not
been well-studied, and we are not aware of any stud-
ies demonstrating a consistent relationship between
the constitutive PRI level and other measures of LUE.
Consequently, caution must be applied when attempting
to use PRI as a LUE indicator, particularly when these
individual factors have not been controlled or dened,
as is the case for most landscape-level remote sensing
studies.
For the facultative component of PRI, irradiance
is particularly critical, in part because the asymptotic
shape of the photosynthetic light-response curve results
in a steadily increasing level of excess radiation and
progressive lowering of LUE as a leaf is exposed to
progressively higher irradiance (Björkman and Dem-
mig-Adams, 1994). Photoprotective pigment levels are
enhanced in response to this excessive irradiance, and
diurnal operation of the xanthophyll cycle causes rapid
PRI changes that scale with LUE (Gamon et al., 1992).
However, this effect on LUE typically relaxes when
leaves are returned to low light (e.g., overnight).
Over longer time scales (days to months), levels of
carotenoids (including carotenes and xanthophylls) can
all change in response to sun exposure during leaf devel-
Gamon and Berry / PRI in sun and shade conifer needles
87
opment (Thayer and Bjorkman, 1990; Demmig-Adams,
1998). Even within a pigment class (e.g., chlorophylls),
the relative levels of pigment types (chl a and b) vary
with sun and shade (Thayer and Björkman, 1990), as
plants adjust the sizes of their photosynthetic light
harvesting complexes relative to their reaction centers
(Bjorkman, 1981). Clearly, pigment levels tend to fol-
low irradiance gradients within the canopy (Niinemets
et al., 2003), which could cause changes in the PRI.
However, to our knowledge, the relative responses of
facultative and constitutive pigment levels to these
within-canopy gradients, and the implications for LUE,
have not been well studied.
In this study, we hypothesized that both facultative
and constitutive changes would occur within a single
canopy due to strong gradients in irradiance, and that
top-canopy leaves would have a higher capacity for
facultative changes characterized by rapid xanthophyll
cycle pigment conversion. Similarly, we expected that
carotenoid–chlorophyll pool sizes would vary with
canopy position and light history, causing reductions
in PRI towards the top of the canopy. We also expected
that the relative levels of individual carotenoid pigments
(α- and β-carotene, lutein, neoxanthin, violaxanthin,
antheraxanthin, and zeaxanthin) would vary due to the
photoprotective and antenna functions ascribed to these
pigments (Thayer and Björkman, 1990). Our goal was
to explore naturally varying responses with leaf posi-
tion and light history in several conifer forests, using
leaf reectance as a rapid probe of both facultative and
constitutive pigment effects on PRI.
MATERIALS AND METHODS
Four conifer species were included in this study: Doug-
las-r (Pseudotsuga menziesii (Mirb.) Franco), western
hemlock (Tsuga heterophylla (Raf.) Sarg.), ponderosa
pine (Pinus ponderosa Douglas ex C. Lawson), and
jack pine (Pinus banksiana, Lamb.). The eld sites in-
cluded a mixed Douglas-r/hemlock stand near Carson,
Washington, USA (Wind River Canopy Crane site),
a ponderosa pine stand at Black Butte, Oregon, USA
(access tower provided by B. Yoder and M. Ryan), and
a jack pine forest near Thompson, Manitoba, Canada
(BOREAS OJP site).
At the jack pine site, top-canopy branches in full sun
and bottom-canopy branches in the shade were accessed
from a scaffolding tower in July 1996 as part of the
BOREAS study (Sellers et al., 1995, 1997). At the Wind
River site, top-canopy branches of three hemlock trees
and three Douglas-r trees in full sun were accessed
by the crane gondola and selected for measurement in
September 1996. In addition, a single shade branch of
a young hemlock tree growing in deep shade was also
selected. The study did not include a “shade” Douglas-
r because no shade branches could be reached from the
forest oor. At the Black Butte ponderosa pine site in
September 1996, four top-canopy “sun” branches were
accessed from a scaffolding tower, and four bottom-
canopy “shade” branches were reached from the ground
(approx. 2 m high).
Branch tips were covered with black cloth bags either
the previous evening or early in the morning (before
sunrise) to maintain needles in their dark state prior
to reectance sampling. Needle reectance was then
measured using a “leaf reectometer” by attaching the
ber optic probe (UNI410, PP Systems, Haverhill MA,
USA) onto the needle with a leaf clip (UNI501, PP Sys-
tems, Haverhill, MA, USA). This ber probe provided
both a white measuring and actinic light (PPFD equal
to full sun), and a path for reected light to reach the
detector (see Gamon and Surfus, 1999, for further de-
tails and instrument description). The leaf clip provided
a xed optical geometry during sampling, which was
essential for accurate and repeatable needle reectance
measurements. To calculate reectance, each leaf scan
was divided by a scan of a white reference (Spectralon,
LabSphere, North Sutton, NH, USA) taken immediately
prior to leaf measurements. The Photochemical Reec-
tance Index (PRI) was calculated as follows:
PRI = (R531 – R570)/(R531 + R570)
where R indicates reectance, and the subscript in-
dicates the wavelength (in nm) (Gamon et al., 1993;
Peñuelas et al., 1995)
By using dark-adapted leaves as a baseline for PRI
measurements, we were able to experimentally separate
constitutive from facultative pigment effects on PRI.
The assumption was that when measuring dark-adapted
leaves, the prior night’s reversion to the dark state (re-
epoxidation to violaxanthin in the dark) minimized the
inuence of diurnal xanthophyll cycle activity on the
measured PRI. Consequently, under these conditions,
PRI variation was primarily due to the constitutive
changes in pigment pool sizes; in subsequent discussion
we refer to this dark-state PRI measurement as “cPRI.”
To examine reectance kinetics indicative of faculta-
tive xanthophyll cycle conversion, dark-adapted leaves
were sampled continuously during several minutes of
high-light exposure, causing rapid transition from the
“dark” to “light” state. The change in reectance dur-
ing this “dark-to-light transition” was expressed as the
change in the PRI (ΔPRI), by subtracting the PRI value
at 10 min of light exposure from the PRI value from the
initial “dark state” (rst scan upon illumination with the
instrument). Similarly, difference spectra (Δ reectance
Israel Journal of Plant Sciences 60 2012
88
spectra) were calculated as the difference between light-
and dark-state reectance, and were used to visualize the
facultative xanthophyll cycle pigment effects on spec-
tral reectance. Previous studies have shown that this
“dark-to-light” method scales closely with the amount
of de-epoxidized xanthophyll cycle pigments formed, as
well as the total pool size of xanthophyll cycle pigments
(violaxanthin + antheraxanthin + zeaxanthin) present in
the leaf (Gamon and Surfus, 1999).
We hypothesized that dark-state PRI values (cPRI,
reecting constitutive effects) would be lower in top-
canopy leaves relative to bottom-canopy leaves, and
that ΔPRI (reecting facultative effects of the xan-
thophyll cycle) would be higher in top-canopy leaves
than bottom-canopy leaves. We also partitioned total
variation in PRI caused by the xanthophyll cycle (dark-
to-light effects on ΔPRI) and due to pool size changes
(canopy position and species effects on cPRI).
A small amount of leaf material was collected
from each leaf species for later pigment analysis via
HPLC according to Thayer and Björkman, 1990.
These samples were from needles of canopy position
adjacent to that of needles sampled for reectance, but
not necessarily from the same needles or branches, the
assumption being that pigment levels for all leaves of a
given age and light environment adjusted similarly to
ambient light levels. Pigment samples were limited to
several needles from 3–4 “representative” branch tips
per species per light environment (sun or shade) and
were generally collected late in the afternoon without
control of prior light conditions. Because the eld sites
were far from laboratory facilities, direct freezing of
leaf samples in the eld was not possible. For these
reasons, direct assessment of xanthophyll cycle pigment
epoxidation state (EPS) was not possible, and EPS data
were not included in the results presented here. The jack
pine samples were maintained on dry ice for three days,
then transferred to a freezer (–80 °C) prior to pigment
extraction. All other samples were stored in an ice chest
followed by refrigeration for several days before pig-
ment extraction could be completed (about 1 week after
initial collection). Because of the difculties in obtain-
ing comparable samples from conifer needles of varying
shapes and thicknesses, pigment pool sizes were rst ex-
pressed on a projected leaf area basis (μmoles m–2), then
expressed relative to chlorophyll content also expressed
on a leaf area basis (μmoles m–2), resulting in unitless
pigment pool size ratios.
RESULTS
Needle reectance spectra were typical of green
leaves (Gausman, 1985), showing the characteristic
green “hump” near 550 nm due to greater chlorophyll
absorption in the blue and red regions, and increased
reectance in the near-infrared due to increased light
scattering (Fig. 1). Spectra from top-canopy sun needles
(Fig. 1A) and bottom-canopy shade needles (Fig. 1B)
varied slightly in the visible region, with shade leaves
appearing slightly darker (lower green reectance peak)
than sun leaves. This difference was more clearly visible
when directly comparing dark-state spectra of sun and
shade needles side-by-side (Fig. 2A), providing direct
visual evidence for constitutive pigment differences in
sun and shade leaves. Examination of the wavelengths
used for PRI calculation (Fig. 2A) suggested that the
contrasting reectance spectra for sun and shade leaves
could easily inuence PRI, as further discussed below.
Upon conversion from the dark to light state, all
leaves showed subtle declines in apparent reectance
(compare solid and dashed lines, Fig. 1). These declines
were more visible when the changes were expressed as
difference spectra (Fig. 2B), representing the facultative
response of the xanthophyll cycle to sudden irradiance.
This sudden dark–light transition is analogous to the
reectance change that leaves undergo in response
to diurnal changes in irradiance (Gamon et al., 1993;
Gamon and Surfus, 1999). These difference spectra
showed a dip centered at 531 nm that was characteristic
of the xanthophyll cycle pigment conversion (Gamon
Fig. 1. Reectance spectra of dark-adapted jack pine needles
in the initial “dark” state and “light” state after ten minutes of
illumination with bright light (approx. 2000 μmol photons m–2
s–1). Panel A: top-canopy sun leaf. Panel B: bottom-canopy
shade leaf.
Gamon and Berry / PRI in sun and shade conifer needles
89
and Surfus, 1999) and a double dip at 685 and 738 nm,
characteristic of chlorophyll uorescence quenching
(Gamon et al., 1990; Gamon and Surfus, 1999). For a
given species, the dip at 531 nm due to the xanthophyll
cycle was deeper for top-canopy sun needles than for
bottom-canopy shade needles (Fig. 2), suggesting larger
xanthophyll pigment pool sizes and greater potential for
diurnal pigment conversion in the leaves from upper
canopy regions.
Upon sudden exposure to high light, all dark-adapted
leaves (regardless of species or canopy position) showed
a rapid (facultative) decline of PRI that continued over
several minutes (representative example shown in
Fig. 3). Typically, at the range of temperatures encoun-
tered during kinetic measurement, this decline reached
an asymptote by 10 min, allowing us to dene a “delta
PRI” (ΔPRI) as the difference between the initial PRI
value measured in the dark state and the nal PRI value
measured at 10 min of light exposure. In Fig. 3, the
offset between the two dark state PRI values illustrates
constitutive effects of pigment pool sizes, and the PRI
kinetics summarized by ΔPRI illustrate the facultative
effects of xanthophyll cycle activity. Collectively, the
results shown in Figs. 1–3 demonstrate that both consti-
tutive and facultative effects on PRI differed for sun and
shade leaves collected from a single tree canopy (see
further discussion below).
Both cPRI (Fig. 4A) and ΔPRI (Fig. 4B) varied
consistently between top-canopy sun needles and bot-
tom-canopy shade needles for all three species where
both canopy positions were sampled (Douglas-r shade
leaves were not available so were not included in this
comparison). The cPRI values were lower for top-
canopy sun leaves than bottom-canopy shade leaves,
suggesting pigment pool size differences. At the same
time, ΔPRI of dark-adapted leaves given sudden light
exposure was higher for sun leaves than shade leaves,
indicating greater capacity for facultative photoconver-
sion of xanthophyll cycle pigments in sun leaves rela-
tive to shade leaves.
Pool sizes for several carotenoid groups (expressed
relative to chlorophyll) varied consistently between sun
and shade leaves for all species sampled (Fig. 5). For ex-
ample, pool sizes of lutein, xanthophyll cycle pigments
(violaxanthin + antheraxanthin + zeaxanthin combined),
and total carotenoid (all carotenes and xanthophylls
combined) were higher for sun leaves than shade leaves
(note that these were expressed as lower pigment ratios
since chlorophyll was the numerator). Similarly, com-
bined xanthophyll levels (violaxanthin, antheraxanthin,
Fig. 2. Top panel: initial dark-state reectance for sun and
shade jack pine needles, showing wavelengths used for PRI
calculation. Bottom panel: difference spectra (“Δ reec-
tance”—light-state reectance minus dark-state reectance)
for single jack pine needles from the sun-exposed, top of the
canopy (“sun leaf,” dotted line) and the shaded bottom of
the canopy (“shade leaf,” solid line). Changing reectance at
531 nm due to xanthophyll cycle activity and the double dip
near 700 nm due to chlorophyll uorescence quenching are
illustrated. Data from Fig. 1.
Fig. 3. Representative PRI kinetics for top-canopy (“sun”)
and bottom-canopy (“shade”) jack-pine needles. In each case,
dark-adapted needles were exposed to full-sun illumination,
reaching a steady state PRI value after approximately 10 min
(derived from data in Figs. 1 and 2). Also illustrated are the
initial, dark-state PRI (cPRI) values and the ΔPRI values for
each needle.
Israel Journal of Plant Sciences 60 2012
90
zeaxanthin, and lutein together) were consistently
higher for sun leaves than shade leaves. This observa-
tion provided independent support for the observations
based on spectral reectance (e.g., dark-state reectance
plots in Fig. 2A, and PRI comparisons presented in Figs.
3 and 4) that constitutive levels of photoprotective carot-
enoid pigments were higher in sun-exposed, top-canopy
leaves relative to shaded leaves deeper in the canopy.
Comparisons of cPRI to pigment pool sizes revealed
several statistically signicant correlations when all
species and treatments (sun and shade) were combined
(Table 1). The cPRI values were strongly related to
the ratios of chlorophyll to several carotenoid pigment
groups, including total carotenoids, lutein, and total
xanthophylls (violaxanthin, antheraxanthin, zeaxanthin,
and lutein combined) (Table 1). Surprisingly, cPRI was
not closely correlated with the pool size of xanthophyll
cycle pigments (violaxanthin + antheraxanthin + zea-
xanthin), unless lutein (another xanthophyll pigment not
directly involved in the xanthophyll cycle) was added to
this pool. These comparisons between cPRI and carot-
enoid pigment levels are illustrated for lutein (Fig. 6A)
and total carotenoids (all carotenes and xanthophylls
combined, Fig. 6B), the two pigment groupings show-
ing the strongest correlations with cPRI (Table 1). These
results suggest that constitutive carotenoid pigment pool
sizes, not xanthophyll cycle pigments, are the primary
source of PRI variability across species and canopy
positions.
To help put these results in context, we compared
absolute values of PRI variation due to facultative causes
(ΔPRI from “dark–light” PRI transitions), and due to
constitutive causes (cPRI variation with canopy posi-
tion and species) (Table 2). Using the facultative PRI
variation (ΔPRI) in top-canopy sun leaves as a reference
(100%), it was clear that shaded leaves had a smaller
(60%) variation in PRI associated with dark-light xan-
thophyll cycle conversion (ΔPRI). On the other hand,
cPRI differences associated with canopy position were
slightly greater (107%) than the range in PRI due to fac-
ultative effects (ΔPRI in top-canopy leaves). When only
top-canopy leaves were considered, cPRI differences
associated with species were almost three times larger
(272%) than facultative effects for the same top-canopy
leaves. When only shade leaves were considered, cPRI
variation due to species was 2.5 times (247%) the facul-
tative effects (ΔPRI) in top-canopy leaves, and 4.5 times
the facultative effects (ΔPRI) of bottom-canopy leaves.
From this analysis, we conclude that total variation in
PRI due to pigment pool sizes associated with canopy
position and species differences were larger than the PRI
variation due to the activity of the xanthophyll cycle.
DISCUSSION
The reectance and Δ reectance spectra for conifer
needles reported here are similar to those reported
previously for broad leaves measured with the same
instrument (Gamon and Surfus, 1999). All species and
sun conditions showed the characteristic feature at
531 nm due to xanthophyll pigment conversion and the
double dip near 700 nm due to chlorophyll uorescence
quenching (Gamon et al., 1990; Gamon and Surfus,
1999). The PRI kinetics upon transition from the dark
state to the light state (full-sun illumination for 10 min)
are similar to patterns reported previously for other
species (Gamon and Surfus, 1999). These similarities
with other reectance spectra and with reectance ki-
Fig. 4. Panel A: Dark-state PRI (cPRI) values for sun and
shade leaves for four conifer species. Panel B: Delta (Δ) PRI
values for sun and shade leaves for the same species (Psme
= Pseudotsuga menziesii, Tshe = Tsuga heterophylla, Pipo =
Pinus ponderosa, and Piba = Pinus banksiana). P. mensiezii
shade leaves were not sampled. Bars and error bars indicate
mean values and standard deviation, respectively.
Gamon and Berry / PRI in sun and shade conifer needles
91
Fig. 5. Pigment ratios for top-canopy (“sun”) and bottom-canopy (“shade”) leaves. Psme = Pseudotsuga menziesii, Tshe = Tsuga
heterophylla, Pipo = Pinus ponderosa, Piba = Pinus banksiana. Chl = total chlorophyll, V = violaxanthin, A = antheraxanthin,
Z = zeaxanthin, L = lutein, Carotenoid = all carotenes and xanthophylls. Bars and error bars indicate mean values and standard
deviation, respectively.
Table 1
R2 values (and p values in parentheses) for linear regressions
between dark-state PRI (cPRI) and the pigment ratios indicated
in the left column. Chl = total chlorophyll, L = lutein,V+A+Z
= total pool size of xanthophyll cycle pigments (violaxanthin,
antherazanthin, and zeaxanthin), V+A+Z+L = total pool size
of xanthophyll cycle pigments plus lutein, Car = all carotenes
and xanthophylls together
Pigment ratios R2 (p)
Chl/(V+A+Z) 0.113 (0.462)
Chl/(V+A+Z+L) 0.7423 (0.013)
Chl/L 0.953 (< 0.001)
Chl/Car 0.879 (0.002)
and xanthophylls combined) in sun leaves versus shade
leaves indicate greater investment in photoprotection
for sun-exposed, top-canopy leaves relative to shade
leaves. This nding is consistent with previous reports
of constitutive pool size differences in sun and shade
leaves (Thayer and Bjorkman, 1990; Demmig-Adams,
1998; Niinemets et al., 2003). The lower cPRI values in
top-canopy sun leaves (relative to shade leaves), along
with the strong correlations with several chl:carotenoid
ratios (Table 1, Fig. 6), demonstrate that leaf-level PRI
within conifer canopies is strongly inuenced by con-
stitutive changes in photoprotective pigment pool sizes.
The weaker correlations between cPRI and xanthophyll
cycle pigment pools (violaxanthin + antheraxanthin +
zeaxanthin) suggest that variation in PRI was not pri-
marily driven by short-term xanthophyll cycle pigment
activity in this case. These results support our hypoth-
esis that natural variation in PRI within plant stands
can be largely explained by constitutive pigment pool
size changes rather than facultative xanthophyll cycle
effects alone.
netics from broadleaved species demonstrate that this
sampling method can be effectively applied to narrow
conifer needles in situ, providing a portable tool for rap-
idly probing pigment light responses for intact conifer
needles in the eld.
The higher levels of carotenoid pool sizes (carotenes
Israel Journal of Plant Sciences 60 2012
92
relation with cPRI is obtained; this correlation was not
signicant with the xanthophyll cycle pigments alone
(Table 1). Similarly, the highest correlation between
cPRI and carotenoid pigment pool sizes was found with
the chl:lutein ratio (Table 1). Recently, a second “xan-
thophyll cycle” involving lutein deepoxidase has been
discovered (Bungard et al., 1999). Like the xanthophyll
cycle involving violaxanthin deepoxidation, the lutein
deepoxidase cycle appears to be involved in energy dis-
sipation (García-Plazaola et al., 2007). As of this writ-
ing, the lutein deepoxidation cycle has been observed in
many species, including some trees (García-Plazaola et
al., 2002; Matsubara et al., 2007), but to our knowledge
its presence or function in conifers has not yet been
conclusively demonstrated. Similarly, the inuence of
lutein deepoxidation on reectance spectra and PRI has
not yet been explored. The high correlation reported
here between cPRI and lutein suggests that lutein may
exert a particular inuence on PRI, perhaps through a
reversible lutein epoxidation. Exploration of this hy-
pothesis was beyond the scope of this initial eld study,
and further experimental work is needed to explore the
possible inuence of lutein deepoxidation on PRI vari-
ability in nature.
In our study, facultative effects associated with xan-
thophyll cycle activity were a smaller source of PRI
variation than constitutive effects; cPRI variation asso-
ciated with canopy position was slightly larger than the
amount of PRI variation (ΔPRI) caused by xanthophyll
cycle activity. When different species were considered,
the cPRI variation was 3–4 times larger than that due to
facultative xanthophyll cycle effects (ΔPRI). In some
ecosystems (e.g., Sims et al., 2006), seasonal variation
in PRI values is much larger than diurnal PRI values.
Together, these observations suggest that PRI variation
in nature is largely driven by constitutive effects, with
facultative effects a smaller component of the total PRI
variation, particularly when all seasons, species, and
stand levels are included in the analysis. The methods
presented here, employing a combination of dark-state
sampling and dark–light conversion, provide one way
to experimentally isolate the two effects in situ without
extensive destructive sampling.
Within each species, the larger ΔPRI values in sun
vs. shade leaves is consistent with the greater invest-
ment in xanthophyll cycle pigment pools (violaxanthin
+ antheraxanthin, + zeaxanthin) in sun-exposed, top-
canopy leaves relative to shade leaves (Figs. 4B and
5A). Consequently, the capacity for short-term changes
in PRI (i.e., facultative effects associated with rapid
xanthophyll pigment conversion) appears to vary with
canopy position in a manner similar to constitutive vari-
ation in pigment pool sizes, a nding that is consistent
Table 2
Variation in PRI attributable to xanthophyll cycle activity
(ΔPRI) or to constitutive pigment changes (cPRI) according
to sun-shade, and species differences. Percent variation is
expressed relative to xanthophyll cycle activity in sun leaves
(100%)
Source of Average PRI %
variation variation variation
Xanthophyll cycle activity in sun 0.0363 100
(ΔPRI)
Xanthophyll cycle activity in shade 0.0219 60
(ΔPRI)
Sun-shade differences (cPRI) 0.0390 107
Species differences, sun leaves (cPRI) 0.0989 272
Species difference, shade leaves (cPRI) 0.0898 247
Fig. 6. Dark-state PRI (cPRI) vs pigment ratios for chl/lu-
tein and chl/carotenoids (all carotenes and xanthophylls
combined). Each point is an average of samples from 3 or 4
branches, and error bars indicate SEM. Data replotted from
Figs. 4 and 5.
Our pigment data (Table 1, Fig. 6) suggest that cPRI
values may be strongly affected by variation in lutein
levels within the canopy. For example, when lutein is
added to the xanthophyll cycle pigments (violaxanthin +
antheraxanthin + zeaxanthin + lutein), a signicant cor-
Gamon and Berry / PRI in sun and shade conifer needles
93
with other eld studies showing correlation between
xanthophyll cycle activity and carotenoid pool sizes
(Stylinki et al., 2002; Filella et al., 2009). Because these
facultative and constitutive effects on PRI are often
strongly interrelated, it is often hard to isolate the short-
er-term xanthophyll effects on PRI from the longer-term
pool size effects on PRI without careful analysis of leaf
pigments or reectance properties, an analysis that is
usually lacking in remote sensing studies of PRI.
Our results have important implications for recent
attempts to apply PRI from aircraft and satellite plat-
forms as a measure of photosynthetic activity. Most
publications on this topic invoke the xanthophyll cycle
as a primary driver of variation in PRI, and most do not
consider constitutive pigment pool size variation with
season, species, or canopy position. Few of these stud-
ies measure actual leaf reectance or pigment levels to
identify the cause of PRI variation, leaving the mecha-
nistic interpretation of remote PRI measurements unre-
solved. Our results suggest that part of the PRI variation
with canopy position (and possibly with sensor view
angle) could be due to constitutive pool size effects on
PRI. Consequently, more work is clearly needed to bet-
ter understand the extent to which constitutive pool size
variation is determining PRI variation when measured
from aircraft or satellite. Without directly considering
various pigment pools, the causes and time scales of
variation in pigment content, and the environmental
context (light history, temperature, moisture, nutrient
conditions, etc.), remote sensing studies employing PRI
from aircraft or satellite platforms could be misconstru-
ing the cause of variation in PRI. Constitutive pigment
effects on PRI could explain some of the scatter in re-
mote sensing studies of the PRI–LUE relationships, and
may partly explain why different vegetation stands do
not always appear to have the same PRI–LUE relation-
ships (Grace et al., 2007; Goerner et al., 2011).
Because few of these recent studies have considered
constitutive pigment effects on photosynthetic activity,
it is unclear to what extent these pool size variations
scale with variation in LUE or photosynthetic activity.
Because these constitutive pool size effects can be larger
than the facultative xanthophyll cycle effects (Table 2),
they can “magnify” the PRI signal, making it appear
more effective as an index of photosynthetic rate or
light-use efciency, particularly if the xanthophyll cycle
and long-term pigment pool size changes are coordi-
nately regulated. We hypothesize that such combined
(facultative and constitutive) pigment effects on PRI
may partly explain why so many remote sensing studies
are able to correlate PRI variation with LUE variation
(Nichol et al., 2000; Rahman et al., 2001, 2004; Drolet
et al., 2005; Garbulsky et al., 2008; Hilker et al., 2009;
Goerner et al., 2011). In this case, variation in pigment
pool sizes makes PRI differences more detectable with
remote sensing. However, due to the different time con-
stants of facultative and constitutive pigment responses,
we might also expect these two types of responses to
inuence light-use efciency over different time scales,
and this can cause disjunct PRI–LUE patterns when
observed over many months or years (Filella et al.,
2004; Sims et al., 2006). The varying contribution of
facultative and constitutive effects could also explain
part of the scatter in many published PRI–LUE rela-
tionships, particularly if the facultative and constitu-
tive components are out of phase. Similarly, we do not
know to what extent the lutein deepoxidase cycle may
contribute to the observed variation in PRI within forest
canopies and further confound (or improve) PRI as a
photosynthetic index. Until these issues are resolved, it
will be difcult to attain the goal of reliably “measuring
photosynthesis from space” (Grace et al., 2007; Coops
et al., 2010; Hall et al., 2011) with PRI. Recent studies
that correct for sunlit fraction (Hall et al., 2008) are an
exciting step in this direction, but may need to consider
additional pigment responses besides the facultative
xanthophyll cycle alone.
Our ndings provide clear evidence for both facul-
tative and constitutive pigment effects on PRI within
a single vegetation canopy. The constitutive pigment
effects explained most of the PRI variation between spe-
cies and within a single canopy. Facultative and consti-
tutive effects operate over different time scales and are
only part of a wide array of mechanisms (physiological
to structural) that plants employ to adjust their radiation
balance under changing environmental and physiologi-
cal conditions. A better appreciation of the full range of
plant responses, from the fast physiological adjustments
to the slower structural and ontogenetic changes, is now
needed to improve our understanding of photosynthetic
physiology from remote sensing. A clearer understand-
ing of both constitutive and facultative changes in leaf
pigment pools will be needed to develop effective
photosynthesis models using PRI. The leaf reectance
methods demonstrated here, along with additional opti-
cal methods for assessing relative pigment levels (e.g.,
Gitelson et al., 2006), could be instrumental in realizing
this goal.
ACKNOWLEDGMENTS
Completion of this study was inspired by the work of
Anatoly Gitelson, who has long been an innovator in as-
sessing leaf pigment content from spectral reectance.
John Scott Surfus provided technical assistance for all
parts of this study. Measurements at the Wind River
Israel Journal of Plant Sciences 60 2012
94
Crane site were conducted with the help of the skilled
staff of the Wind River Crane facility. Canopy access
and sampling advice at the Black Butte site was pro-
vided by Barbara Bond and Mike Ryan. Measurements
at the BOREAS Old Jack Pine site were conducted with
eld assistance by Art Fredeen. Financial support for
this work was provided by grants from the US-EPA,
NASA, and NSF to J. Gamon and a NASA BOREAS
grant to J. Berry.
REFERENCES
Björkman, O. 1981. Responses to different quantum ux den-
sities. In: Lange, O.L., Nobel, P.S., Osmond, C.B., Ziegler,
H., eds. Encyclopedia of plant physiology, N.S. Vol. 12A:
Physiological plant ecology—interactions with the physi-
cal environment. Springer, Heidleberg, pp. 57–107.
Björkman, O., Demmig-Adams, B. 1994. Regulation of photo-
synthetic light energy capture, conversion and dissipation
in leaves of higher plants. In: Schulze, E.-D., Caldwell,
M.M., eds. Ecophysiology of photosynthesis. Ecological
Studies, Vol. 100. Springer Verlag, Berlin, pp. 17–47.
Brugnoli, E., Björkman, O. 1992. Chloroplast movement in
leaves: inuence on chlorophyll uorescence and measure-
ments of light-induced absorbance changes related to ΔpH
and zeaxanthin formation. Photosynth. Res. 32: 23–35.
Bungard, R.A., Ruban, A.V., Hibberd, J.M., Press, M.C., Hor-
ton, P., Scholes, J.D. 1999. Unusual carotenoid composi-
tion and a new type of xanthophyll cycle in plants. Proc.
Natl. Acad. Sci. USA 96: 1135–1139.
Coops, N.C., Hilker, T., Hall, F.G., Nichol, C.J., Drolet, G.G.
2010. Estimation of light-use efciency of terrestrial
ecosystems from space: a status report. BioScience 60:
788–797.
Demmig-Adams, B. 1998. Survey of thermal energy dissipa-
tion and pigment composition in sun and shade leaves.
Plant Cell Physiol. 39: 474–482.
Demmig-Adams, B., Adams, W.W.III. 1992. Photoprotection
and other responses of plants to high light stress. Annu.
Rev. Plant Physiol. Plant Mol. Biol. 43: 599–626.
Drolet, G.G., Huemmrich, K.F., Hall, F.G., Middleton, E.M.,
Black, T.A., Barr, A.G., Margolis, H.A. 2005. A MODIS-
derived photochemical reectance index to detect inter-an-
nual variations in the photosynthetic light-use efciency
of a boreal deciduous forest. Remote Sens. Environ. 98:
212–224.
Filella, I., Peñuelas, J. Llorens, L., Estiarte, M. 2004. Re-
ectance assessment of seasonal and annual changes in
biomass and CO2 uptake of a Mediterranean shrubland
submitted to experimental warming and drought. Remote
Sens. Environ. 90: 308–318.
Filella, I., Porcar-Castell, A., Munné- Bosch, S., Bäck, J., Gar-
bulsky, M., Peñuelas, J. 2009. PRI assessment of long-term
changes in carotenoids/chlorophyll ratio and short-term
changes in de-epoxidation state of the xanthophyll cycle.
Int. J. Remote Sens. 30: 4443–4455.
Gamon, J.A., Pearcy, R.W. 1989. Leaf movement, stress
avoidance and photosynthesis in Vitis californica. Oecolo-
gia 79: 475–481.
Gamon, J.A., Surfus, J.S. 1999. Assessing leaf pigment con-
tent and activity with a reectometer. New Phytologist
143: 105–117.
Gamon, J.A., Field, C.B., Bilger, W., Björkman, O., Fredeen,
A., Peñuelas, J. 1990. Remote sensing of the xanthophyll
cycle and chlorophyll uorescence in sunower leaves and
canopies. Oecologia 85: 1–7.
Gamon, J.A., Peñuelas, J., Field, C.B. 1992. A narrow-wave-
band spectral index that tracks diurnal changes in photo-
synthetic efciency. Remote Sens. Environ. 41: 35–44.
Gamon, J.A., Filella, I., Peñuelas, J. 1993. The dynamic
531-nanometer ∆ reectance signal: a survey of twenty
angiosperm species. In: Yamamoto, H.Y., Smith, C.M.,
eds. Photosynthetic responses to the environment. Ameri-
can Society of Plant Physiologists, Rockville, MD, pp.
172–177.
Gamon, J.A., Serrano, L., Surfus, J.S. 1997. The photochemi-
cal reectance index: an optical indicator of photosynthetic
radiation-use efciency across species, functional types,
and nutrient levels. Oecologia 112: 492–501.
Gamon, J.A., Field, C.B., Fredeen, A.L., Thayer, S. 2001.
Assessing photosynthetic downregulation in sunower
stands with an optically-based model. Photosynth. Res.
67: 113–125.
Garbulsky, M.F., Peñuelas, J., Papale, D., Filella, I. 2008.
Remote estimation of carbon dioxide uptake by a Mediter-
ranean forest. Global Change Biol. 14: 2860–2867.
Garcia-Plazaola, J.I., Faria, T., Abadia, J., Chaves, M.M.,
Pereira, J.S. 1997. Seasonal changes in xanthophyll com-
position of cork oak (Quercus suber L.) leaves under Medi-
terranean climate. J. Exp. Bot. 48: 1667–1674.
García-Plazaola, J.I., Hernández, A., Errasti, I., Becerril, J.M.
2002. Occurrence and operation of the lutein epoxide cycle
in Quercus species. Funct. Plant Biol. 29: 1075–1080.
García-Plazaola, J.I., Matsubara, S., Osmond, C.B. 2007. The
lutein epoxide cycle in higher plants: its relationships to
other xanthophyll cycles and possible functions. Funct.
Plant Biol. 34: 759–773.
Gausman, H.W. 1985. Plant leaf optical properties. Texas Tech
Univ. Press, Lubbock, TX, USA.
Gitelson, AA., Merzlyak, M.N. 1994. Spectral reectance
changes associated with autumn senescence of Aesculus-
hippocastanum L. and Acer platanoides L. leaves—spec-
tral features and relation to chlorophyll estimation. J. Plant
Physiol. 143: 286–292.
Gitelson, A.A., Keydan, P., Merzlyak, M.N. 2006. Three-
band model for noninvasive estimation of chlorophyll,
carotenoids, and anthocyanin contents in higher plant
leaves. Geophys. Res. Lett. 33: L11402, doi:10.1029/
2006GL026457.
Gitelson, A.A., Chivkunova, B., Merzlyak, N. 2009. Nonde-
structive estimation of anthocyanins and chlorophylls in
anthocyanic leaves. Am. J. Bot. 96: 1861–1868.
Goerner, A., Reichstein, M., Tomelleri, E., Hanan, N., Ram-
bal, S., Papale, D., Dragoni, D., Schmullius, C. 2011.
Gamon and Berry / PRI in sun and shade conifer needles
95
Remote sensing of ecosystem light use efciency with
MODIS-based PRI. Biogeosci. 8: 189–202.
Grace, J., Nichol, C., Disney, M., Lewis, P., Quaife, T., Bower,
P. 2007. Can we measure terrestrial photosynthesis from
space directly using spectral reectance and uorescence?
Global Change Biol. 13: 1484–1497.
Hall, F.G., Hilker, T., Coops, N.C., Luyapustin, A.,
Huemmrich, K.F., Middleton, E., Margolis, H., Drolet,
G., Black, A.T. 2008. Multi-angle remote sensing of forest
light use efciency by observing PRI variation with canopy
shadow fraction. Remote Sens. Environ. 112: 3201–3211.
Hall, F.G., Hilker, T., Coops, N.C. 2011. PHOTOSYNSAT,
photosynthesis from space: theoretical foundations of a
satellite concept and validation from tower and spaceborne
data. Remote Sens. Environ. 115: 1918–1925.
Hilker, T., Lyapustin, A., Hall, F.G., Wang, Y., Coops, N.C.,
Drolet, G., Black, T.A. 2009. An assessment of photosyn-
thetic light use efciency from space: modeling the atmo-
spheric and directional impacts on PRI reectance. Remote
Sens. Environ. 113: 2463–2475.
Jiang, C.-D., Gao, H.-Y., Zou, Q., Jiang, G.-M., Li, L.-H.
2006. Leaf orientation, photorespiration and xanthophyll
cycle protect young soybean leaves against high irradiance
in eld. Environ. Exp. Bot. 55: 87–96.
Matsubara, S., Morosinotto, T., Osmond, C.B., Bassi, R. 2007.
Short- and long-term operation of the lutein-epoxide cycle
in light-harvesting antenna complexes. Plant Physiol. 144:
926–941.
Merzlyak, M.N., Gitelson, A.A. 1995. Why and what for the
leaves are yellow in autumn—on the interpretation of opti-
cal-spectra of senescing leaves (Acer platanoides L). J.
Plant Physiol. 145: 315–320.
Merzlyak, M.N., Gitelson, A.A., Chivkunova, O.B., Rakitin,
V.Y. 1999. Non-destructive optical detection of pigment
changes during leaf senescence and fruit ripening. Physiol.
Plant. 106: 135–141.
Merzlyak, M.N., Gitelson, A.A., Chivkunova, O.B., So-
lovchenko, A.E., Pogosyan, I. 2003. Application of reec-
tance spectroscopy for analysis of higher plant pigments.
Russian J. Plant Physiol. 50: 704–710.
Nichol, C.J., Huemmrich, K.F., Black, T.A., Jarvis, P.G.,
Walthall, C.L., Grace, J., Hall, F.G. 2000. Remote sens-
ing of photosynthetic-light-use efciency of boreal forest.
Agri. Forest Meteo. 101: 131–142.
Niinemets, U., Kollist, H., Garzia-Plazaola, J.I., Hernandez,
A., Becerril, J.M. 2003. Do the capacity and kinetics for
modication of xanthophyll cycle pool size depend on
growth irradiance in temperate trees? Plant Cell Environ.
26: 1787–1801.
Park, Y.I., Chow, W.S., Anderson, J.M. 1996. Chloroplast
movement in the shade plant Tradescantia albiora helps
protect photosystem II against light stress. Plant Physiol.
111: 867–875.
Peñuelas, J., Filella, I., Gamon, J.A. 1995. Assessment of
photosynthetic radiation-use efciency with spectral re-
ectance. New Phytol. 131: 291–296.
Rahman, A.F., Gamon, J.A., Fuentes, D.A., Roberts, D.A.,
Prentiss, D. 2001. Modeling spatially distributed eco-
system ux of boreal forests using hyperspectral indices
from AVIRIS imagery. J. Geophys. Res. 106(D24):
33,579–33,591.
Rahman, A.F., Cordova, V.D., Gamon, J.A., Schmid, H.P.,
Sims, D.A. 2004. Potential of MODIS ocean bands for
estimating CO2 ux from terrestrial vegetation: A novel
approach. Geophys. Res. Lett. 31: L10503, doi:10.1029/
2004GL019778.
Sellers, P., Hall, F., Margolis, H., Kelly, B., Baldocchi, D., den
Hartog, G., Cihlar, J., Ryan, M. G., Goodison, B., Crill,
P., Ranson, J., Lettenmaier, D., Wickland, D.E. 1995. The
voreal Ecosystem–Atmosphere Study (BOREAS): An
overview and early results from the 1994 eld year. Bull.
Am. Meteo. Soc. 76: 1549–1577.
Sellers, P., Hall, F., Kelly, R., Black, A., Baldocchi, D., Berry,
J., Ryan, M., Ranson, K.J., Crill, P., Lettenmaier, D., Mar-
golis, H., Cihlar, J., Newcomer, J., Fitzjarrald, D., Jarvis,
P., Gower, S.T., Halliwell, D., Williams, D., Goodison, B.,
Wickland, D., Guertin, F. 1997. BOREAS in 1997: experi-
ment overview, scientic results, and future directions. J.
Geophys. Res. 102(D24): 28731–28769.
Sims, D.A., Gamon, J.A. 2002. Relationships between leaf
pigment content and spectral reectance across a wide
range of species, leaf structures and developmental stages.
Remote Sens. Environ. 81: 337–354.
Sims, D.A., Luo, H., Hastings, S., Oechel, W.C., Rahman,
A.F., Gamon, J.A. 2006. Parallel adjustments in vegeta-
tion greenness and ecosystem CO2 exchange in response
to drought in a Southern California chaparral ecosystem.
Remote Sens. Environ. 103: 289–303.
Steele, M.R., Gitelson, A.A., Rundquist, D.C., Merzlyak,
M.N. 2009. Nondestructive estimation of anthocy-
anin content in grapevine leaves. Am. J. Enol. Vitic. 60:
87–92.
Stylinski, C.D., Gamon, J.A., Oechel, W.C. 2002. Seasonal
patterns of reectance indices, carotenoid pigments and
photosynthesis of evergreen chaparral species. Oecologia
131: 366–374.
Thayer, S.S., Björkman, O. 1990. Leaf xanthophyll content
and composition in sun and shade determined by HPLC.
Photosynth. Res. 23: 331–343.
Ustin, S.L., Gitelson, A.A., Jacquemoud, S., Schaepman,
M.E., Asner, G.P., Gamon, J.A., Zarco-Tejada, P. 2009. Re-
trieval of foliar information about plant pigment systems
from high resolution spectroscopy. Remote Sens. Environ.
113: S67–77.
Zygielbaum, A.I., Arkebauer, T.J., Walter-Shea, E.A., Scoby,
D.L. 2012. Detection and measurement of vegetation pho-
toprotection stress response using PAR reectance. Isr. J.
Plant Sci. 60: 37–47 (this issue).
... The vegetative index PRI is sensitive to rapid changes (within a day) in the state of de-epoxidation, which is a signal of the mutual transformation of xanthophylls. This vegetative index can also be used to track long-term changes in the ratio of chlorophyll and carotenoid pools [45,54,[63][64][65][66]. PRI indirectly reflects the water content of vegetation [67][68][69][70]. ...
Article
Full-text available
Conifers are a common type of plant used in ornamental horticulture. The prompt diagnosis of the phenological state of coniferous plants using remote sensing is crucial for forecasting the consequences of extreme weather events. This is the first study to identify the “Vegetation” and “Dormancy” states in coniferous plants by analyzing their annual time series of spectral characteristics. The study analyzed Platycladus orientalis, Thuja occidentalis and T. plicata using time series values of 81 vegetation indices and 125 spectral bands. Linear discriminant analysis (LDA) was used to identify “Vegetation” and “Dormancy” states. The model contained three to four independent variables and achieved a high level of correctness (92.3 to 96.1%) and test accuracy (92.1 to 96.0%). The LDA model assigns the highest weight to vegetation indices that are sensitive to photosynthetic pigments, such as the photochemical reflectance index (PRI), normalized PRI (PRI_norm), the ratio of PRI to coloration index 2 (PRI/CI2), and derivative index 2 (D2). The random forest method also diagnoses the “Vegetation” and “Dormancy” states with high accuracy (97.3%). The vegetation indices chlorophyll/carotenoid index (CCI), PRI, PRI_norm and PRI/CI2 contribute the most to the mean decrease accuracy and mean decrease Gini. Diagnosing the phenological state of conifers throughout the annual cycle will allow for the effective planning of management measures in conifer plantations.
... Two facets of plant response to stress that reflect the rapid change in plant absorption and the slower, seasonal, or ontogenetic pigment pool size adjustments have been termed the 'facultative' and 'constitutive' responses, respectively (Gamon and Berry 2012). Constitutive changes in plant absorption based on detection of chlorophyll content or carotenoid-to-chlorophyll ratio at leaf or plant level are well developed; examples of which are presented in the introduction. ...
Article
Full-text available
Environmental stresses on plants are among the major limitations to crop productivity. Plants are naturally equipped with a set of acclimatory mechanisms enabling them to withstand stresses without irreversible damage. Here, we overhauled the approach to non-invasive gauging of the response of maize to combined drought and highlight stresses. This was done to leverage the advantages of a recently developed framework based on using the reflected signal absorption coefficient modality. We document the deployment of two responses, one based on non-photochemical quenching (NPQ) of the absorbed light energy and the second on the adjustment of leaf photosynthetically active radiation (PAR) interception. The NPQ-based response, implemented by the xanthophyll cycle and inferred from changes in the reflection of light in the blue-green region of the spectrum, engaged at the beginning of the stress but quickly reached a plateau. We demonstrate that altering leaf absorption in PAR is a fundamental plant response mechanism to diverse stresses. This response is quantified by the PAR absorption coefficient retrieved from the non-destructively measured reflectance. A profound decrease in the PAR absorption coefficient in the whole PAR region was shown to be a reliable measure of the degree of stress regardless of its cause. The effects of superficial shading and leaf position were readily detectable by the leaf PAR absorption coefficient. We consider this report a proof of concept for quantifying plant stress and monitoring the acclima-tion state via the absorption coefficient in the PAR region through non-destructive, remotely sensed techniques. ARTICLE HISTORY
... In addition, the results of the study confirm the need for remote spectral monitoring of vegetation, taking into account its light history, which was previously pointed out by other authors [38][39][40][41]. ...
Article
Full-text available
Hyperspectral imaging techniques are widely used to remotely assess the vegetation and physiological condition of plants. Usually, such studies are carried out without taking into account the light history of the objects (for example, direct sunlight or light scattered by clouds), including light-stress conditions (photoinhibition). In addition, strong photoinhibitory lighting itself can cause stress. Until now, it is unknown how light history influences the physiologically meaningful spectral indices of reflected light. In the present work, shifts in the spectral reflectance characteristics of Ficus elastica leaves caused by 10 h exposure to photoinhibitory white LED light, 200 μmol photons m−2 s−1 (light stress), and moderate natural light, 50 μmol photons m−2 s−1 (shade) are compared to dark-adapted plants. Measurements were performed with a Cubert UHD-185 hyperspectral camera in discrete spectral bands centred on wavelengths from 450 to 950 nm with a 4 nm step. It was shown that light stress leads to an increase in reflection in the range of 522–594 nm and a decrease in reflection at 666–682 nm. The physiological causes of the observed spectral shifts are discussed. Based on empirical data, the light-stress index (LSI) = mean(R666:682)/mean(R552:594) was calculated and tested. The data obtained suggest the possibility of identifying plant light stress using spectral sensors that remotely fix passive reflection with the need to take light history into account when analysing hyperspectral data.
... The relationship between PRI and NPQ is complicated at the leaf scale, where their correlation is influenced not only by PRI observation error but also by non-qE components in NPQ, although several studies have built empirical regression relationships between these two variables [16,17]. Previous studies have classified PRI as "facultative PRI" and "constitutive PRI" according to its timescale response [27,51]. For leaves under high illumination, PRI presented a stronger relationship with the qE component than the non-qE component, which elucidated the finding of Sukhova and Sukhov [52], who showed the connection between ∆PRI (modified formation) and the qE component. ...
Article
Full-text available
Non-photochemical quenching (NPQ) is an indicator of crop stress. Until now, only a limited number of studies have focused on how to estimate NPQ using remote sensing technology. The main challenge is the complicated regulatory mechanism of NPQ. NPQ can be divided into ener-gy-dependent (qE) and non-energy-dependent (non-qE) quenching. The contribution of these two components varies with environmental factors, such as light intensity and stress level due to the different response mechanisms. This study aims to explore the feasibility of estimating NPQ using photosynthesis-related vegetation parameters available from remote sensing by considering the two components of NPQ. We concurrently measured passive vegetation reflectance spectra by spec-trometer, as well as active fluorescence parameters by pulse-amplitude modulated (PAM) of rice (Oryza sativa) leaves. Subsequently, we explored the ability of the selected vegetation parameters (including the photochemical reflectance index (PRI), inverted red-edge chlorophyll index (IRECI), near-infrared reflectance of vegetation (NIRv), and fluorescence quantum yield (ΦF)) to estimate NPQ. Based on different combinations of these remote sensing parameters, empirical models were established to estimate NPQ using the linear regression method. Experimental analysis shows that the contribution of qE and non-qE components varied under different illumination conditions. Under high illumination, the NPQ was attributed primarily to the qE component, while under low illumination, it was equally attributed to the qE and non-qE components. Among all tested pa-rameters, ΦF was sensitive to the qE component variation, while IRECI and NIRv were sensitive to the non-qE component variation. Under high illumination, integrating ΦF in the regression model captured NPQ variations well (R2 > 0.74). Under low illumination, ΦF, IRECI, and NIRv explained 24%, 62%, and 65% of the variation in NPQ, respectively, while coupling IRECI or NIRv with ΦF considerably improved the accuracy of NPQ estimation (R2 > 0.9). For all the samples under both low and high illumination, the combination of ΦF with at least one of the other parameters (in-cluding IRECI, NIRv and PAR) offers a more versatile and reliable approach to estimating NPQ than using any single parameter alone. The findings of this study contribute to the further development of remote sensing methods for NPQ estimation at the canopy scale in the future.
... For evergreens, zeaxanthin and other carotenoid levels remain high during cold period, maximizing dissipation of light energy to protect leaves during winter (Adams III et al., 2002;Öquist & Huner, 2003). In this case, PRI variation over long time periods is influenced by the seasonal change in leaf pigment pool sizes (Gamon & Berry, 2012;Garbulsky et al., 2011;Hmimina et al., 2014;Wong & Gamon, 2015a, 2015b, which are closely tied to seasonal gross primary productivity (GPP) patterns (Wong et al., 2020). Consequently, PRI can indicate both short-term (e.g., diurnal) downregulation due to xanthophyll cycle pigment activity, and longer-term (e.g., seasonal) adjustments in photosynthetic activity due to adjustments in photosynthetic and photoprotective pigment pools. ...
Article
Full-text available
Located at northern latitudes and subject to large seasonal temperature fluctuations, boreal forests are sensitive to the changing climate, with evidence for both increasing and decreasing productivity, depending upon conditions. Optical remote sensing of vegetation indices based on spectral reflectance offers a means of monitoring vegetation photosynthetic activity and provides a powerful tool for observing how boreal forests respond to changing environmental conditions. Reflectance-based remotely sensed optical signals at northern latitude or high-altitude regions are readily confounded by snow coverage, hampering applications of satellite-based vegetation indices in tracking vegetation productivity at large scales. Unraveling the effects of snow can be challenging from satellite data, particularly when validation data are lacking. In this study, we established an experimental system in Alberta, Canada including six boreal tree species, both evergreen and deciduous, to evaluate the confounding effects of snow on three vegetation indices: the normalized difference vegetation index (NDVI), the photochemical reflectance index (PRI), and the chlorophyll/carotenoid index (CCI), all used in tracking vegetation productivity for boreal forests. Our results revealed substantial impacts of snow on canopy reflectance and vegetation indices, expressed as increased albedo, decreased NDVI values and increased PRI and CCI values. These effects varied among species and functional groups (evergreen and deciduous) and different vegetation indices were affected differently, indicating contradictory, confounding effects of snow on these indices. In addition to snow effects, we evaluated the contribution of deciduous trees to vegetation indices in mixed stands of evergreen and deciduous species, which contribute to the observed relationship between greenness-based indices and ecosystem productivity of many evergreen-dominated forests that contain a deciduous component. Our results demonstrate confounding and interacting effects of snow and vegetation type on vegetation indices and illustrate the importance of explicitly considering snow effects in any global-scale photosynthesis monitoring efforts using remotely sensed vegetation indices.
... In support of this finding, the PRI index of leaf reflectance was high at pre-dawn and midday measurements in both DCAP and HP plants (Fig. 8). PRI was originally developed to track the xanthophyll cycle pigment changes and whereby the light use efficiency (Levizou & Manetas, 2007;Gamon & Berry, 2012;Vanikiotis, Stagakis & Kyparissis, 2021), but has also been related to carotenoid to chlorophyll ratio in green leaves (Sims & Gamon, 2002). Notably, PRI variation over the diurnal cycle is primarily considered a function of xanthophyll cycle changes in response to stress, whilst long-term changes in PRI (during weeks or months) mainly reflect modifications of car-to-chl ratio as an acclimation to light environment or stress (Sims & Gamon, 2002). ...
Article
Full-text available
Background Aquaponics is an innovative farming system that combines hydroponics and aquaculture, resulting in the production of both crops and fish. Decoupled aquaponics is a new approach introduced in aquaponics research for the elimination of certain system bottlenecks, specifically targeting the optimization of crops and fish production conditions. The aquaponics-related literature predominantly examines the system’s effects on crop productivity, largely overlooking the plant functional responses which underlie growth and yield performance. The aim of the study was the integrated evaluation of basil performance cultivated under coupled and decoupled aquaponic systems compared with a hydroponic one, in terms of growth and functional parameters in a pilot-scale aquaponics greenhouse. Methods We focused on the efficiency of the photosynthetic process and the state of the photosynthetic machinery, assessed by instantaneous gas exchange measurements as well as photosynthetic light response curves, and in vivo chlorophyll a fluorescence. Light use efficiency was estimated through leaf reflectance determination. Photosynthetic pigments content and leaf nutritional state assessments completed the picture of basil functional responses to the three different treatments/systems. The plant’s functional parameters were assessed at 15-day intervals. The experiment lasted for two months and included an intermediate and a final harvest during which several basil growth parameters were determined. Results Coupled aquaponics resulted in reduced growth, which was mainly ascribed to sub-sufficient leaf nutrient levels, a fact that triggered a series of negative feedbacks on all aspects of their photosynthetic performance. These plants experienced a down-regulation of PSII activity as reflected in the significant decreases of quantum yield and efficiency of electron transport, along with decreased photosynthetic pigments content. On the contrary, decoupled aquaponics favored both growth and photochemistry leading to higher light use efficiency compared with coupled system and hydroponics, yet without significant differences from the latter. Photosynthetic light curves indicated constantly higher photosynthetic capacity of the decoupled aquaponics-treated basil, while also enhanced pigment concentrations were evident. Basil functional responses to the three tested production systems provided insights on the underlying mechanisms of plant performance highlighting key-points for systems optimization. We propose decoupled aquaponics as an effective system that may replace hydroponics supporting high crops productivity. We suggest that future works should focus on the mechanisms involved in crop and fish species function, the elucidation of which would greatly contribute to the optimization of the aquaponics productivity.
Article
Anthropic potentially toxic element (PTE) releases can lead to persistent pollution in soil. Monitoring PTEs by their detection and quantification on large scale is of great interest. The vegetation exposed to PTEs can exhibit a reduction of physiological activities, structural damage … Such vegetation trait changes impact the spectral signature in the reflective domain 0.4-2.5 μm. The objective of this study is to characterize the impact of PTEs on the spectral signature of two pine species (Aleppo and Stone pines) in the reflective domain and ensure their assessment. The study focuses on nine PTEs: As, Cr, Cu, Fe, Mn, Mo, Ni, Pb, Zn. The spectra are measured by an in-field spectrometer and an aerial hyperspectral instrument on a former ore processing site. They are completed by measurements related to vegetation traits at needle and tree scales (photosynthetic pigments, dry matter, morphometry …) to define the most sensitive vegetation parameter to each PTE in soil. A result of this study is that chlorophylls and carotenoids are the most correlated to PTE contents. Context-specific spectral indices are specified and used to assess metal contents in soil by regression. These new vegetation indices are compared at needle and canopy scales to literature indices. Most of the PTE contents are predicted at both scales with Pearson correlation scores between 0.6 and 0.9, depending on species and scale.
Article
Full-text available
The anthocyanin (Anth) content in leaves provides valuable information about the physiological status of plants. Thus, there is a need for accurate, efficient, practical methodologies to estimate this biochemical parameter. Reflectance measurement is a means of quickly and nondestructively assessing leaf Anth content in situ. The objective of this study was to test the overall performance and accuracy of nondestructive techniques for estimating Anth content in grapevine leaves. Relationships were established between Anth content and four vegetation indices: NIR (near-infrared)/green, red/green, anthocyanin reflectance index (ARI, based on reflectances in bands within the green and the red-edge regions), and a modified anthocyanin reflectance index (MARI, based on reflectances in green, red edge, and NIR). The algorithms for Anth retrieval were calibrated. The accuracy of Anth prediction was evaluated using an independent data set containing sampled leaves from two field-grown grape cultivars (Saint Croix and Saint Pepin) with no adjustment of the coefficients after initial calibration. Although Anth in the validation data set was widely variable, from 3 to 45 nmol cm-2, the ARI and MARI algorithms were capable of accurately predicting Anth content in grapevine leaves with a root mean square error below 3 nmol cm-2 and 2.3 nmol cm-2, respectively. Such an approach has potential for developing simple hand-held field instrumentation for accurate nondestructive Anth estimation and for analyzing digital airborne or satellite imagery to assist in making informed decisions regarding vineyard management.
Article
Full-text available
A critical variable in the estimation of gross primary production of terrestrial ecosystems is light-use efficiency (LUE), a value that represents the actual efficiency of a plant's use of absorbed radiation energy to produce biomass. Light-use efficiency is driven by the most limiting of a number of environmental stress factors that reduce plants' photosynthetic capacity; these include short-term stressors, such as photoinhibition, as well as longer-term stressors, such as soil water and temperature. Modeling LUE from remote sensing is governed largely by the biochemical composition of plant foliage, with the past decade seeing important theoretical and modeling advances for understanding the role of these stresses on LUE. In this article we provide a summary of the tower-, aircraft-, and satellite-based research undertaken to date, and discuss the broader scalability of these methods, concluding with recommendations for ongoing research possibilities.
Article
Optical reflectance from leaf surfaces has been known for decades to increase with decreasing leaf water content. Experimental results show this increase consistently in maize in the visible (photosynthetically active radiation—PAR) and middle infrared (MIR) spectral regions, and with weaker correlation in the near infrared (NIR) region. Changes in chlorophyll concentration have been shown to be too small to substantially contribute to increasing reflectance during the duration of these experiments. Therefore, the reflectance responses to water deficit are perplexing since reflectance in the PAR region is dominated by pigment absorption, while reflectance in the MIR region is dominated by water molecule absorption. We report on recent experiments with maize that indicate that the reflectance changes during water stress in the PAR and MIR regions are due to changing optical absorption, whilst those in the NIR region are due to changing optical scatter. In addition, reflectance in PAR and MIR appears to be influenced by down-welling irradiance. These findings are consistent with the idea that increasing PAR reflectance induced by water stress may result, primarily, from chloroplast avoidance movement, a plant photoprotective response.
Article
A new xanthophyll cycle involving de-epoxidation of lutein epoxide into lutein in the light and epoxidation back in the dark has been recently described in parasitic plants. In the present work, the presence of the non-ubiquitous lutein epoxide was detected in many different non-parasitic woody plant species. Phylogenetic constraints are critical for the presence of this xanthophyll, since the largest amounts are found within the Fagaceae family and mainly in the genus Quercus. Irrespective of their ecological characteristics, this xanthophyll was found in eight Quercus species. Under photoinhibitory conditions lutein epoxide decreased in parallel with violaxanthin, and a concomitant increase in lutein was observed, indicating operation of this cycle in the genus Quercus. However, recovery in darkness differed from that of the xanthophyll cycle. It is also shown that lutein epoxide content is higher in shade leaves, especially during summer. The operation of this cycle could also contribute to photoprotection mechanisms of evergreen Quercus species during winter.
Article
As a part of our investigations to test the hypothesis that zeaxanthin formed by reversible de-epoxidation of violaxanthin serves to dissipate any excessive and potentially harmful excitation energy we determined the influence of light climate on the size of the xanthophyll cycle pool (violaxanthin + antheraxanthin + zeaxanthin) in leaves of a number of species of higher plants. The maximum amount of zeaxanthin that can be formed by de-epoxidation of violaxanthin and antheraxanthin is determined by the pool size of the xanthophyll cycle. To quantitate the individual leaf carotenoids a rapid, sensitive and accurate HPLC method was developed using a non-endcapped Zorbax ODS column, giving baseline separation of lutein and zeaxanthin as well as of other carotenoids and Chl a and b. The size of the xanthophyll cycle pool, both on a basis of light-intercepting leaf area and of light-harvesting chlorophyll, was ca. four times greater in sun-grown leaves of a group of ten sun tolerant species than in shade-grown leaves in a group of nine shade tolerant species. In contrast there were no marked or consistent differences between the two groups in the content of the other major leaf xanthophylls, lutein and neoxanthin. Also, in each of four species examined the xanthophyll pool size increased with an increase in the amount of light available during leaf development whereas there was little change in the content of the other xanthophylls. However, the α-carotene/β-carotene ratio decreased and little or no α-carotene was detected in sun-grown leaves. Among shade-grown leaves the α-carotene/β-carotene ratio was considerably higher in species deemed to be umbrophilic than in species deemed to be heliophilic. The percentage of the xanthophyll cycle pool present as violaxanthin (di-epoxy-zeaxanthin) at solar noon was 96–100% for shade-grown plants and 4–53% for sun-grown plants with zeaxanthin accounting for most of the balance. The percentage of zeaxanthin in leaves exposed to midday solar radiation was higher in those with low than in those with high photosynthetic capacity. The results are consistent with the hypothesis that the xanthophyll cycle is involved in the regulation of energy dissipation in the pigment bed, thereby preventing a buildup of excessive excitation energy at the reaction centers.
Article
The reflectance, transmittance and absorption spectra of Acer platanoides leaves were recorded in the progress of full-term autumn senescence and compared with absorbance spectra of extracts from, and pigment contents in, these leaves. The observed changes in leaf spectra and development of intensive yellow color of the leaves are considered as related to the changes in the light depth penetration in the photosynthetic tissues accompanying chlorophyll breakdown. It is suggested that carotenoids, serving as effective light traps (in spite on their relatively low concentrations), are able to provide protection against harmful effects of blue-light irradiation; it can explain the physiological significance of their retention in leaves up to terminal stages of the senescence process.