Download Expression and Inhibition of the Carboxylating

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Tissue engineering wikipedia , lookup

Cell encapsulation wikipedia , lookup

Cellular differentiation wikipedia , lookup

Organ-on-a-chip wikipedia , lookup

Cell culture wikipedia , lookup

JADE1 wikipedia , lookup

List of types of proteins wikipedia , lookup

Amitosis wikipedia , lookup

Photosynthesis wikipedia , lookup

Transcript
Expression and Inhibition of the Carboxylating and
Decarboxylating Enzymes in the Photosynthetic
C4 Pathway of Marine Diatoms1
Patrick J. McGinn* and Franc
xois M.M. Morel
Department of Geosciences, Princeton University, Princeton, New Jersey 08540
There is evidence that the CO2-concentrating mechanism in the marine diatom Thalassiosira weissflogii operates as a type of singlecell C4 photosynthesis. In quantitative-PCR assays, we observed 2- to 4-fold up-regulation of two phosphoenolpyruvate carboxylase (PEPC) gene transcripts in Thalassiosira pseudonana cells adapted to low pCO2, but did not detect such regulation in
Phaeodactylum tricornutum grown under similar conditions. Transcripts encoding phosphoenolpyruvate carboxykinase did
not appear to be regulated by pCO2 in either diatom. In T. pseudonana and T. weissflogii, net CO2 fixation was blocked by 3,
3-dichloro-2-(dihydroxyphosphinoyl-methyl)-propenoate (a specific inhibitor of PEPC), but was restored by about 50% and 80%,
respectively, by addition of millimolar concentrations of KHCO3. In T. pseudonana, T. weissflogii, and P. tricornutum, rates of net
O2 evolution were reduced by an average of 67%, 55%, and 62%, respectively, in the presence of 20 mM quercetin, also an inhibitor of
PEPC. Quercetin promoted net CO2 leakage from inhibited cells to levels in excess of the equilibrium CO2 concentration,
suggesting that a fraction of the HCO32 taken up is fated to leak back into the medium as CO2 when PEPC activity is blocked. In
parallel to these experiments, in vitro assays on crude extracts of T. pseudonana demonstrated mean inhibition of 65% of PEPC
activity by quercetin. In the presence of 5 mM 3-mercaptopicolinic acid (3-MPA), a classic inhibitor of phosphoenolpyruvate
carboxykinase, photosynthetic O2 evolution was reduced by 90% in T. pseudonana. In T. weissflogii and P. tricornutum, 5 mM 3-MPA
totally inhibited net CO2 fixation and O2 evolution. Neither quercetin nor 3-MPA had a significant inhibitory effect on
photosynthetic O2 evolution or CO2 uptake in the marine chlorophyte isolates Chlamydomonas sp. or Dunaliella tertiolecta. Our
evidence supports the idea that C4-based CO2-concentrating mechanisms are generally distributed in diatoms. This conclusion is
discussed within the context of the evolutionary success of diatoms.
Diatoms fix approximately 1015 g of CO2 into organic
carbon every year, equivalent to roughly 40% of marine
primary production (Granum et al., 2005). Although
these phytoplanktons play a pivotal role in mediating
the exchange of inorganic carbon between the air-sea
interface and the deep ocean (Falkowski and Raven,
1997), our understanding of how they effect this transformation is incomplete. The low affinity of algal Rubisco
for CO2 and its propensity to fix O2 instead of CO2 in
an oxygenase reaction (photorespiration; Kaplan and
Reinhold, 1999) impose a chronic limitation on the rate
of CO2 fixation. In seawater, the concentration of inorganic carbon is high (approximately 2 mM), but most of
it is HCO32, whereas only CO2 can serve as the substrate for Rubisco. The concentration of CO2 is around
10 mM (more in upwelling regions and less during
blooms; Granum et al., 2005), lower than estimates of
the half-saturation constant (KmCO2) for diatom Rubisco,
which are in the range of 40 to 60 mM (Badger et al., 1998;
1
This work was supported by the National Science Foundation
(grant no. 0351499 to F.M.M.M.).
* Corresponding author; e-mail [email protected].
The author responsible for distribution of materials integral to
the findings presented in this article in accordance with the policy
described in the Instructions for Authors (www.plantphysiol.org) is:
Patrick J. McGinn ([email protected]).
www.plantphysiol.org/cgi/doi/10.1104/pp.107.110569
300
Tortell, 2000). Clearly, the concentration of CO2 in seawater should limit the rate of carbon fixation.
To ameliorate this problem, many eukaryotic phytoplankton species have evolved an energy-dependent
CO2-concentrating mechanism (CCM), induced when
cells are exposed to low pCO2. For example, in chlorophytes, HCO32 is accumulated in the chloroplast and a
high CO2 supply rate to Rubisco is ensured by the
rapid dehydration of HCO32 catalyzed by a carbonic
anhydrase (CA; EC 4.2.1.1) colocalized with Rubisco
(Kaplan and Reinhold, 1999; Fig. 1A). An external CA
expressed under low pCO2 maintains a high rate of CO2
supply to the whole cell by catalyzing the dehydration
of the large pool of extracellular HCO32 (Moroney et al.,
1989).
Available data strongly suggest that the CCM of diatoms is biochemically similar to the CCM of C4 plants
(Fig. 1B). Indeed, pioneering work on the path of carbon in marine phytoplankton revealed stark differences
between diatoms and green microalgae, with the earliest products of fixation in diatoms typically identified as C4 and amino acids (Beardall and Morris, 1975;
Beardall et al., 1976). Instead of being transported into
the chloroplast, the HCO32 taken up from the medium
is fixed into the four-carbon organic acid oxaloacetate
(OAA) by phosphoenolpyruvate (PEP) carboxylase
(PEPC; EC 4.1.1.31) in the cytoplasm. In the marine diatom Thalassiosira weissflogii, PEPC activity was shown
to increase during growth under low pCO2 or when
Plant Physiology, January 2008, Vol. 146, pp. 300–309, www.plantphysiol.org Ó 2007 American Society of Plant Biologists
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2008 American Society of Plant Biologists. All rights reserved.
Carbon-Concentrating Mechanisms in Diatoms
Figure 1. Current models of CCM function in C. reinhardtii and other eukaryotic phytoplankton (A) and in marine diatoms (B).
Mechanistic details are given in the text.
CA activity was limited by preconditioning at low zinc
or by short-term treatment with the CA inhibitor ethoxyzolamide, suggesting that PEPC activity was necessary under low pCO2 or when the cell’s ability to express
CA was compromised (Reinfelder et al., 2000). A specific inhibitor of PEPC, the PEP analog 3,3-dichloro2-dihydroxyphosphinoylmethyl-2-propenoate (DECP;
Jenkins et al., 1987) was shown to potently inhibit photosynthesis in T. weissflogii. Under high CO2:O2 ratios
(which discourage photorespiration), this compound
showed no inhibitory effect. It was also ineffective in a
marine Chlamydomonas isolate, which utilizes a wellcharacterized C3 carbon assimilation pathway (Reinfelder
et al., 2004). The product of PEPC, OAA, has been shown
to support O2 evolution in the chloroplast and both OAA
and malate (reduced form of OAA) have been shown
to inhibit 14CO2 fixation, but not O2 evolution when
applied exogenously to T. weissflogii cells (Reinfelder
et al., 2004). Short-term pulse-chase experiments have
shown significant initial 14C labeling in malate followed by transfer of label to phosphosugars, such as
3-phosphoglyceric acid in T. weissflogii, which is the
characteristic pattern of transfer in C4 plants (Reinfelder
et al., 2000; Morel et al., 2002). More recently, short-term
14
C-labeling experiments have confirmed the incorporation of inorganic carbon into malate in T. weissflogii,
but not in Thalassiosira pseudonana (Roberts et al., 2007).
It is possible that the short-term C4 labeling in T. pseudonana might have occurred more rapidly than could
be measured in the first 2 s of this study. Alternatively,
the difference between these two diatoms could represent genuine intraspecific variations in the type of
CCM employed. Whereas evidence for C4 photosynthesis in diatoms is accumulating, it is clear that these
organisms maintain one important component of conventional CCMs—that of active uptake of HCO32, which
is essential for rapid photosynthesis in the marine en-
vironment (Tortell et al., 1997). Indeed, in terms of primary inorganic acquisition, recent work has shown
that diatoms acclimate to low pCO2 by increasing their
affinity for both CO2 and HCO32 and that HCO32
transport contributes greater than one-half the carbon
required for photosynthesis and growth (Burkhardt et al.,
2001; Rost et al., 2003).
In our model for the CCM of diatoms, the OAA is
transported by a dedicated dicarboxylic acid transport
system into the chloroplast where it is subsequently
decarboxylated by a second enzyme, PEP carboxykinase (PCKase; EC 4.1.1.49). The CO2 released by this
reaction is subsequently captured through the operation of a typical Calvin cycle in the pyrenoid structure
of the chloroplast and the regenerated PEP is translocated back to the cytoplasm (Reinfelder et al., 2000).
One mole of ATP is required for decarboxylation of
1 mol of OAA to CO2 and PEP. Reinfelder et al. (2000)
found that PCKase copurified with Rubisco in cell fractionation experiments, suggesting chloroplast localization. These authors also measured PCKase activity
in crude extracts of T. weissflogii and found it sufficient
to support the observed rates of CO2 fixation by Rubisco.
In the marine diatom Skeletonema costatum and the kelp
Laminaria setchellii, PCKase activity has also been found
to localize to the chloroplast, suggesting PCKase may
be commonly found in this compartment in marine
plants (Cabello-Pasini et al., 2001). The ratio of PEPC to
Rubisco activities has been measured in pure diatom
cultures and in field samples (Tortell et al., 2006; Cassar
and Laws, 2007). The small measured values showing
a large excess of Rubisco over PEPC activities have been
put forth as evidence against the existence of a C4 pathway in diatoms. We note that in vitro rates are notoriously difficult to relate to in vivo kinetics, partly because
they measure maximal potential rates and partly because we do not know the degree of enzymatic satu-
Plant Physiol. Vol. 146, 2008
301
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2008 American Society of Plant Biologists. All rights reserved.
McGinn and Morel
Figure 2. Effect of growth pCO2 on steadystate levels of PEPC and PCKase transcripts
(designated PEPC 1, PEPC 2, and PCKase,
respectively) in T. pseudonana (A) and P.
tricornutum (B). Annotated coding sequences were obtained from the Joint Genome Institute (http://www.jgi.doe.gov).
Procedures for total RNA isolation, cDNA
synthesis, and quantitative-PCR amplification
and mRNA detection are given in ‘‘Materials and Methods.’’ Values reported represent the means 6 SD of three (A) or two (B)
separate cultures.
ration in the cell. These data should therefore not be
considered conclusive evidence against the presence
of C4-based CCMs in diatoms.
As in chlorophytes, CAs play a critical role in the
CCM of diatoms (Fig. 1B). Cytoplasmic CA maintains
a low concentration of CO2 in this compartment by
continuously hydrating it to HCO32, which is fixed by
PEPC. This serves to augment the inward diffusive
flux of CO2 from the outside of the cell and to scavenge
unfixed CO2 effluxing from the chloroplast. The inward gradient of CO2 is further augmented by extracellular CA, which maintains CO2 at equilibrium with
HCO32 at the cell surface and prevents the development of an external CO2 diffusion gradient. Morel et al.
(2002) used immunofluorescence techniques to demonstrate that the intracellular CA enzyme is principally cytoplasmic in T. weissflogii and largely absent
from the chloroplasts.
Whether the carbon taken up during photosynthesis
is stored as an organic or inorganic intermediate is a
key difference between C4-based CCMs and more typical CCMs, which accumulate inorganic carbon against
a concentration gradient (Kaplan and Reinhold, 1999;
Badger and Spalding, 2000). Reinfelder et al. (2004)
have addressed this issue and showed in T. weissflogii
that only a small fraction of the carbon assimilated by
the CCM existed as inorganic carbon when photosynthesis was stopped by treatment with a potent killing
solution, which effectively quenched all enzymatic
activity, suggesting that inorganic carbon is efficiently
trapped into a stable organic compound in the cell. In
contrast, in Chlamydomonas suspensions treated with
the same killing solution, a large fraction of the carbon
was released in inorganic form consistent with the
accumulation of HCO32 in the light during photosynthesis.
In this study, we tested several key features of the
model C4-CCM shown in Figure 1B. We examined by
quantitative-PCR the expression in selected diatom
species of the two enzymes PEPC and PCKase thought
to be responsible for the carboxylation and decarboxylation of the C4 intermediate. With the aid of a membrane inlet mass spectrometer (MIMS) system, we also
explored the effects on photosynthetic gas exchange of
inhibitors known to block these enzymes in other organisms. When appropriate, we compared the results
obtained with model diatom species to those obtained
with chlorophytes.
Figure 3. Effect of 3-MPA on steady-state net CO2 uptake and O2
evolution in low pCO2 T. pseudonana cells measured by MIMS. Results
are shown from one representative experiment of two separate analyses.
Cells (0.5–1.8 3 107 cells mL21) were concentrated by filtration and
resuspended in 1 mL of assay buffer (10 mM HEPES, pH 7.5) containing
3.5% (w/v) NaCl. 3-MPA was added from a stock solution (77 mM)
dissolved in 25 mM HEPES (pH 7.5) to a final concentration of 5 mM.
Darkened cells were preincubated in the presence of 3-MPA for 5 min.
Photosynthesis was initiated by illuminating the cell suspension with
white light obtained from a tungsten lamp at an intensity of 300 mmol
photons m22 s21. The residual dissolved inorganic carbon in the assay
buffer (approximately 300 mM) was used as the source of carbon for CO2
fixation.
302
Plant Physiol. Vol. 146, 2008
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2008 American Society of Plant Biologists. All rights reserved.
Carbon-Concentrating Mechanisms in Diatoms
RESULTS AND DISCUSSION
Regulation of PEPC and PCKase Gene Expression
by pCO2
Inspection of the annotated genome sequences of
T. pseudonana and Phaeodactylum tricornutum revealed
the presence of two genes encoding PEPC and a single
PCKase gene in each diatom (http://genome.jgi-psf.
org/euk_home.html). To quantify the effect of pCO2 on
the abundance of PEPC and PCKase gene transcripts,
we conducted quantitative reverse transcription-PCR
analyses of RNA isolated from these two organisms. All
gene expression levels are reported as mRNA copy
number of the gene of interest per copy of the nonregulated actin housekeeper gene. In T. pseudonana, transcripts for each of the two genes apparently encoding
PEPC 1 and PEPC 2 were found to be about 20% as
abundant as the actin under high pCO2 and increased
by a factor of about 3 to about 60% of actin under low
pCO2 (Fig. 2A). In contrast, we found no evidence for
regulation by pCO2 of the homologous genes in P.
tricornutum. However, the absolute abundance of the
PEPC transcripts was high in this diatom, remaining
between 40% to 60% of actin regardless of pCO2 (Fig.
2B). The abundance of the PCKase transcript did not
respond significantly to growth pCO2 in either diatom,
although the transcript levels were again constitutively
much higher in P. tricornutum than in T. pseudonana (Fig.
2, A and B). Similar responses of PEPC and PCKase
gene transcripts to pCO2 in T. pseudonana have recently
been reported by Roberts et al. (2007). These authors
also demonstrated strong diurnal effects on gene expression, with the large subunit of Rubisco down-
regulated at night, whereas the two PEPC isoforms
appeared to be mildly up-regulated.
Differing transcriptional responses of the PEPC genes
to pCO2 could be related to different modes of regulation operating in the two organisms. Apparently, neither of the PEPC enzymes in either diatom is regulated
by posttranslational processes because none contain
the characteristic N-terminal Ser residue that typically
serves as the site of phosphorylation (data not shown).
It thus appears that in T. pseudonana the increase in
abundance of PEPC at low pCO2 potentially represents
transcriptional up-regulation of PEPC activity, although it must be borne in mind that transcript levels
are often an unreliable proxy for the amounts of
corresponding functional enzymes (Gibon et al., 2004).
The apparent lack of regulation of PCKase transcription in this organism may simply reflect the downstream position of this enzyme in the CCM. Regulation
of PEPC as a function of CO2 may result in a relatively
constant supply of C4 organic acid, obviating the need
to regulate the decarboxylating enzyme. In P. tricornutum, the high abundance of PEPC and PCKase transcripts may correspond to a strategy of maintaining
constitutively high capacity to process inorganic carbon.
Sensitivity of Photosynthesis to PCKase Inhibition
in Diatoms
To gain better insight into the importance of PCKase
in the CCM of diatoms, we tested the effects of a specific inhibitor, 3-mercaptopicolinic acid (3-MPA), on
photosynthesis in low pCO2 cells. This compound has
been used as a potent inhibitor of the PCKase enzyme
Figure 4. Effect of 5 mM 3-MPA on steady-state net O2 evolution in the light by low pCO2 cells of T. pseudonana (A), T.
weissflogii (B), P. tricornutum (C), Chlamydomonas sp. (D), and D. tertiolecta (E). Assay conditions were as described for
Figure 3. Cells were concentrated to 18 to 23 3 106 mL21, 1.7 to 2.4 3 106 mL21, 20 to 30 3 106 mL21, 5 to 12 3 106 mL21, and
8.5 3 106 mL21, respectively, by either filtration or centrifugation (see text for details). Rates of photosynthesis were obtained
in the presence/absence of inhibitor in assay medium (10 mM HEPES, pH 7.5, 3.5% NaCl) equilibrated with 400 mM KHCO3 in
300 mmol photons m22 s21. Except for P. tricornutum, all experiments were conducted using two to five independently grown
cultures conditioned to low pCO2 by harvesting the cells at pH . 8.6. Error bars represent the SD about the mean of two to three
independent determinations.
Plant Physiol. Vol. 146, 2008
303
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2008 American Society of Plant Biologists. All rights reserved.
McGinn and Morel
Figure 5. Effect of DCDP on photosynthesis in T. pseudonana (A) and T. weissflogii
(B). Cells were preconcentrated by filtration
and resuspended in assay buffer (10 mM
HEPES, pH 7.6, 3.5% NaCl) in the chamber
of the mass spectrometer and incubated for
a few minutes in the dark. Contaminating
levels of inorganic carbon (about 400 mM)
were used to drive net CO2 fixation. The
light was turned on and, after steady-state
net O2 evolution was obtained, DCDP was
added to a final concentration of 490 mM.
from leaves and isolated chloroplasts of numerous C4
terrestrial and aquatic plants (Ray and Black, 1976, 1977;
Rathnam and Edwards, 1977; Reiskind and Bowes,
1991) and in T. weissflogii (Reinfelder et al., 2004) and is
usually considered to be specific for this enzyme. In
experiments with T. pseudonana, there was little inhibition of photosynthesis until the 3-MPA concentration
reached about 4 mM, above which there was a sharp
decline in CO2 uptake and O2 evolution (Fig. 3). Similar patterns of inhibition of O2 evolution have been
noted in Udotea flabellum, an aquatic C4 macrophyte,
although PCKase functions in a carboxylating, rather
than decarboxylating, role in this plant (Reiskind and
Bowes, 1991). This demonstrates the inhibitory effects
of 3-MPA on steady-state photosynthesis in this marine diatom.
We chose a concentration of 5 mM 3-MPA to further
evaluate its effects on photosynthesis in four other
phytoplankton species, T. weissflogii, P. tricornutum,
Chlamydomonas sp., and Dunaliella tertiolecta (Fig. 4,
A–E). In treated suspensions of T. weissflogii and P. tricornutum, we observed total inhibition of net O2 evolution and CO2 uptake and fixation in the light compared
to inhibition by an average of 90% in T. pseudonana.
Figure 6. Effect of quercetin on net O2 and CO2 exchange during steady-state photosynthesis in cell suspensions of T. pseudonana (A and B) and T. weissflogii (C) measured by MIMS. Cells were concentrated 50- to 100-fold by filtration and resuspended in 1 mL of assay buffer (10 mM HEPES, pH 7.6, 3.5% [w/v] NaCl). A 10-mL aliquot of the cell concentrate was sampled for
estimates of cell density just prior to placing the suspension in the sample cuvette (thermoregulated to 20°C) attached to the mass
spectrometer. Cells were incubated in the dark for approximately 2 min before 20 mM AZ was added to inhibit extracellular CA
activity. Dissolved inorganic carbon for photosynthesis was added as KHCO3 to 500 mM. Light (300 mmol photons m22 s21) was
provided from a tungsten lamp from one side of the cuvette. Photosynthesis was initiated by illuminating the cells. After steadystate CO2 fixation was under way (approximately 10 min), quercetin was added in the light to the indicated concentrations. The
long dashed lines represent the maximal potential CO2 concentration above [CO2 eq] that could be supported by the observed
rate of respiratory CO2 release. AZ stock solutions were maintained at 10 mM in DMSO. In B, 2 mM KHCO3 was added to the
inhibited T. pseudonana suspension (shown in A) approximately 15 min after the initial addition of inhibitor. The double arrow
indicates the suspension was darkened. The numbers in parentheses beside the oxygen traces are the rates of net O2 evolution in
units of mmol O2/108 cells/h.
304
Plant Physiol. Vol. 146, 2008
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2008 American Society of Plant Biologists. All rights reserved.
Carbon-Concentrating Mechanisms in Diatoms
Indeed, in T. weissflogii and P. tricornutum, only net respiratory activity was observed. On average, the rates
of O2 uptake were equivalent to approximately 20% of
the rates of O2 evolution in untreated cells. Contrary to
what we observed in the diatoms, we detected no significant effect of 3-MPA on the rates of photosynthesis
in either Chlamydomonas sp. or D. tertiolecta suspensions.
These results indicate a critical role in CO2 fixation for
PCKase in diatoms, but not in chlorophytes. According to our model of the CCM in diatoms (Fig. 1B),
inhibition by 3-MPA arises from the blockage of OAA
decarboxylation in the chloroplast, which, in turn,
starves Rubisco for CO2 in the C3 Calvin cycle.
Sensitivity of Photosynthesis and PEPC Activity to
DCDP and Quercetin
Previous work has shown a strong effect of DCDP, a
specific inhibitor of PEPC, on CO2 fixation in T. weissflogii and in C4 land plants that can be largely overcome by adding CO2 to high concentration (Jenkins
et al., 1987, 1989; Reinfelder et al., 2004). We revisited
this by testing the effects of this compound on T. pseudonana, which had not been previously shown to be
sensitive to it. At a concentration of 490 mM, net O2
evolution (as a proxy for CO2 fixation) was totally
inhibited in T. pseudonana, but was partially restored
(45%) by the addition of 5 mM KHCO3 (Fig. 5A). T.
weissflogii was also inhibited by DCDP at this concentration, but this was reversed by about 80% when 2 mM
KHCO3 was added (Fig. 5B). Our results with T. weissflogii are in good agreement with Reinfelder et al. (2004)
and, taken together, support the conclusion that T. pseudonana has a similar type of CCM to T. weissflogii.
The flavonoid compound quercetin has previously
been used as a potent inhibitor of PEPC activity in
extracts of both a C4 plant, maize (Zea mays), and a C3
plant, canola (Brassica napus; Pairoba et al., 1996; Moraes
and Plaxton, 2000). We explored the potential for this
compound to inhibit CCM function in phytoplankton
by monitoring the rates of inorganic carbon uptake and
O2 evolution in concentrated cell suspensions. Although,
as an inhibitor of PEPC, quercetin does not discriminate between the enzymes of C4 and C3 plants, its
effect on photosynthesis should be relatively more
severe in phytoplankton employing a C4-CCM compared to those utilizing a conventional CCM based on
inorganic carbon accumulation.
We examined the effect of quercetin on photosynthesis and CO2 exchange in actively photosynthesizing
T. pseudonana and T. weissflogii cells (Fig. 6, A and C,
respectively; Table I). Within 1 min of illumination
(started at time 0), the cells drew [CO2] below the
concentration at equilibrium with HCO32, [CO2 eq],
indicating a net CO2 influx into the cells. We believe this
influx results from passive CO2 diffusion across the
plasmalemma, down a concentration gradient set up by
the activities of CA and PEPC in the cytoplasm (Fig. 1B).
In T. pseudonana suspensions, addition of 20 mM quercetin inhibited the rate of O2 evolution by 80% to 90%
almost immediately, and it subsequently recovered to
about 20% of the preinhibited rate (Fig. 6A). The [CO2]
in the medium rose rapidly to a value roughly 15%
above [CO2 eq] and remained at this level for the
ensuing 10 min of the experiment. In T. weissflogii
suspensions, after an initial 20 mM quercetin addition,
the [CO2] rose in the medium slowly and surpassed
[CO2 eq] after about 7 min (Fig. 6C). Addition of a
second 20 mM pulse of quercetin resulted in 85% inhibition of net O2 evolution and caused the [CO2] to rise
more rapidly and to a higher level than the first pulse.
The requirement in this particular experiment for twice
the amount of inhibitor to elicit the same response as in
T. pseudonana may have been due to the comparatively
larger amount of biomass used (2.43 3 109 mm3 cell
volume/mL for T. weissflogii versus 0.9 3 109 mm3 cell
volume/mL for T. pseudonana). Quantitative analysis of
the CO2 fluxes in the presence of the inhibitor (see
‘‘Materials and Methods’’) revealed that only 65% and
50% of the observed excess of [CO2] over [CO2 eq] could
be accounted for by respiratory CO2 release in T.
Figure 7. Effect of quercetin on apparent PEPC activity in low CO2 cells
of T. pseudonana. Cells were ground in a chilled mortar and pestle for
two periods of 7 min in extraction buffer (50 mM BICINE, pH 7.5, 1.5 M
glycerol, 10 mM MgCl2, 1 mM EDTA, 10 mM NaHCO3, 5 mM dithiothreitol, and 5 mg mL21 bovine serum albumin). Crude extracts were preincubated at room temperature for 45 min, at which time they were
subdivided for treatment. To one vial, 1 mCi of H14CO32 was added and,
to a second vial, 1 mCi of H14CO32 was added along with quercetin to a
final concentration of 0.2 mM. These reactions were initiated by adding
5 mM PEP followed by a 2-h incubation at 25°C. A third vial was
included, which received radiolabel but did not receive PEP. These
treatments acted as negative controls. Reactions were terminated by
transferring aliquots to vials containing HCl (35%, approximately 9 N).
Acidified extracts were evaporated to dryness and 14C incorporated into
organic carbon was counted by liquid scintillation.
Plant Physiol. Vol. 146, 2008
305
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2008 American Society of Plant Biologists. All rights reserved.
McGinn and Morel
Figure 8. Effect of quercetin on photosynthetic CO2 and O2 exchange in concentrated cell suspensions of Chlamydomonas sp.
(A) and D. tertiolecta (B). Assay conditions were as described for Figure 6. Quercetin was added to a final concentration of 40 mM
for both phytoplankton species.
pseudonana and T. weissflogii suspensions, respectively
(Fig. 6, A and C, long dashed line). Analyses of several
such experiments in both diatoms produced essentially
the same result. The most parsimonious explanation for
the excess CO2 efflux is that it originates from the pool
of cytoplasmic HCO32. Bicarbonate is still taken up by
the cells after quercetin addition as shown by the
residual O2 evolution; part of that HCO32 leaks back
into the medium as CO2 after dehydration by the
cytoplasmic CA. Both the inhibition of photosynthesis
resulting from the inhibition of PEPC and the simultaneous CO2 leakage are fully consistent with our model
of the diatom CCM (Fig. 1B). Similar to the DCDP
experiments (Fig. 5), addition of a high concentration of
inorganic carbon partially reversed the inhibitory effect
of quercetin, pointing to an inhibition of the cell’s
ability to acquire carbon, rather than a nonspecific
effect on photosynthesis (Fig. 6B). In three separate
experiments with quercetin, addition of 2 mM KHCO3
restored the rate of photosynthesis by an average of
42% 6 5%. In T. weissflogii cells inhibited with quercetin,
the rate of photosynthesis was restored by about 38%
by 2 mM KHCO3 addition (data not shown).
The observed effects of quercetin on diatoms provide powerful support to the view that PEPC mediates
HCO32 fixation in the cytoplasm and that it constitutes
the major pathway for carbon fixation in diatoms if
indeed quercetin acts by inhibiting PEPC activity. Because quercetin has been shown to inhibit various enzymes and processes in cells (e.g. Lang and Racker,
1974), we performed two types of controls: in vitro
assays with cell extracts and in vivo assays with chlorophytes that are known to operate a classic C3 photosynthetic pathway. We also examined the effects of
flavonoids other than quercetin on photosynthesis in
T. pseudonana.
We tested the sensitivity of diatom PEPC to quercetin, with in vitro activity assays in crude extracts of
low-CO2-grown T. pseudonana cells. In two separate
assays on independent cultures, we detected a mean
inhibition of PEPC activity of approximately 65% with
0.2 mM quercetin (Fig. 7). The concentration of quercetin required for inhibition of PEPC activity in crude
extracts (0.2 mM) was larger than that necessary for
inhibition of in vivo photosynthesis (20–40 mM). This
difference is likely due to the much higher concentration of HCO32 present in the assay medium (10 mM)
compared to the in vivo gas exchange experiments
(0.5–1.5 mM HCO32) because HCO32 has been shown
to decrease the sensitivity of some PEPC isoforms to
inhibitors (Parvathi et al., 1998). Although the biochemical mechanism of quercetin inhibition of PEPC is
not known with certainty, Pairoba et al. (1996) reported
competitive inhibition with respect to PEP in the enzyme from maize leaves.
We explored the potential for quercetin to inhibit
CCM function in Chlamydomonas sp. and D. tertiolecta.
At a concentration of 40 mM quercetin, there was no
apparent effect on either CO2 uptake or O2 evolution in
either of these phytoplankton species (Fig. 8, A and B).
At a concentration of 200 mM quercetin, only a small
effect on the rates of photosynthetic gas exchange was
detected in Chlamydomonas sp. cells (data not shown).
The contrasting effects of quercetin on the diatoms
T. pseudonana, T. weissflogii, and P. tricornutum (inhibited
Table I. Effect of quercetin, baicalein, and rutin hydrate on
steady-state net O2 evolution in low CO2 cells of T. pseudonana
and their associated apparent Ki values for PEPC inhibition
Rates of photosynthesis are reported in units of mmol O2/108 cells/h.
*, Values from Pairoba et al. (1996).
Flavonoid
Quercetin
Baicalein
Rutin hydrate
Net O2 Evolution
Control
25 mM
75 mM
10.4 6 1.7
9.8 6 1.5
9.4
3.4 6 1.3
1.4 6 1.6
9.3
n.d.
n.d.
7.7
306
Ki (mM)*
2.9
6.5
19.8
Plant Physiol. Vol. 146, 2008
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2008 American Society of Plant Biologists. All rights reserved.
Carbon-Concentrating Mechanisms in Diatoms
Table II. Effect of quercetin on the rate of photosynthesis in
selected marine phytoplankton
Rates are reported in units of mmol O2/108 cells/h. Chlamydomonas
sp. and D. tertiolecta were treated with 40 mM quercetin. All other
treatments were 20 mM. Results obtained with T. pseudonana are
presented in Table I. Experimental details are given in the legends to
Figures 6 and 8. Values are the mean of two to three independent
measurements on separate cultures. Error 5 1 SD about the mean.
Phytoplankton Species
Control Rate
1Quercetin
T. weissflogii
P. tricornutum
Chlamydomonas sp.
D. tertiolecta
39
4.5
53.9
18.1
18
1.6
55.5
17.4
6
6
6
6
2.1
0.8
3.4
1.8
6
6
6
6
12
0.7
4.5
0.9
by 67%, 55%, and 62%, respectively, at 20 mM quercetin)
and the chlorophytes Chlamydomonas sp. and D. tertiolecta (no inhibition) are shown in Table II. The effect of
quercetin on photosynthesis and CO2 fluxes observed
in diatoms is not seen in C3 phytoplankton. This is
consistent with the inhibition of PEPC, which should
cripple photosynthesis in organisms with a C4-based
CCM but have little or no effect on C3 organisms.
We evaluated the effects of two additional flavonoid
compounds, baicalein and rutin hydrate, on CCM
function in T. pseudonana (Table I). Baicalein appeared
to be as effective an inhibitor of photosynthesis as
quercetin, inhibiting O2 evolution by about 85% at a
concentration of 25 mM. Conversely, rutin hydrate did
not inhibit photosynthesis at all at this concentration.
In the presence of 75 mM rutin hydrate, only a modest
inhibition of about 20% was detected (Table I). These
results are wholly consistent with the relative effectiveness of flavonoids as inhibitors of PEPC as documented
from the Ki values obtained from in vitro enzymatic
assays (Pairoba et al., 1996; Table I). The consistency
between the relative effects of flavonoids on photosynthesis in diatoms and their effectiveness as PEPC inhibitors provides further support to the notion that the
former is caused by the latter.
vantages might they have gained by adopting a C4
pathway for inorganic carbon acquisition? Obviously,
the O2-insensitive fixation of carbon by PEPC can minimize photorespiratory activity in marine phytoplankton just as well as in terrestrial C4 plants. A C4-based
CCM is also particularly efficient for maintaining low
CO2 in the cytoplasm while ensuring an efficient rate
of fixation by Rubisco. The primary fixation of HCO32
into OAA effectively traps the carbon in a form with
low potential for efflux back into the medium. Indeed,
the low carbon isotope discrimination factor of diatoms is evidence for a low rate of CO2 leakage during
steady-state photosynthesis. In other words, diatoms
appear to be among the low pCO2 specialists (Clark and
Flynn, 2000) and this may be the key to their global
ecological success: A C4-based CCM might make them
well adapted to the low pCO2 of the present-day atmosphere and surface waters to which they have contributed through their efficient export of carbon to the
deep sea.
Another possible benefit a C4-CCM in diatoms is to
integrate efficiently carbon and nitrogen metabolism
(Allen et al., 2006). Plants and microalgae carboxylate
PEP via PEPC in an anaplerotic reaction to replenish
carbon-rich substrates into the TCA cycle to support
anabolic metabolism. For example, Guy et al. (1989) and
Elrifi and Turpin (1986) demonstrated the diversion of
carbon away from the Calvin cycle by the anaplerotic
reaction to provide carbon skeletons for amino acid
synthesis in nitrogen-limited Selenastrum minutum fed
with a transient pulse of NO32 or NH41. Most regions of
the oceans are nitrogen limited and phytoplankton
capable of rapidly exploiting new inputs of nitrogen
can potentially out-compete other groups. Maintaining
high PEPC activity could poise diatoms in a physiological state, whereby newly fixed carbon could either
be diverted toward supporting amino acid synthesis
during transitions from nitrogen deficiency to nitrogen
replenishment or could be used as the primary path of
carbon fixation in the chloroplast when nitrogen is not
limiting.
C4-CCMs: Biochemical and Ecological Perspectives
Our results provide additional support for the operation of a C4-based CCM in diatoms. These protists
appeared some 200 to 300 millions years ago and are
now among the most successful photosynthetic autotrophs in freshwater and oceans. What competitive ad-
CONCLUSION
Taken at face value, our data show a striking difference between the carbon acquisition systems of diatoms and chlorophytes: The former are inhibited by
Table III. Oligonucleotide primers used in quantitative-PCR analyses in this study
Species
Encoded Gene
Forward Primer (5#/3#)
Reverse Primer (5#/3#)
T. pseudonana
PEPC 1
PEPC 2
PCKase
Actin
PEPC 1
PEPC 2
PCKase
Actin
AGATGTATGAACAGTGGCCCTGGT
GGAGTTGGCGAGGCTATCAATGAA
ACTCGAGAAACACAGTGCTAACGC
ACTGGATTGGAGATGGATGG
CGAAAGTCTTCGCGCTATTC
CGGTTTGTTCCCTACTTTCG
TCAGCCACATTCTCGACTTG
TGACAGAGCGTGGTTACTCG
AGAGGACGCTCGCATCAATTGAAC
GTCGCGCACCAAAACATTCTCA
TCGGGATTGGACCAGGTACTCTTT
CAAAGCCGTAATCTCCTTCG
ACCATTTCGACCAAATCCAC
TTCGCTTTGATCTGTCGTTG
TACGCTTCCCGGTACCATAG
ACCATCCATCTCCAAACCAC
P. tricornutum
Plant Physiol. Vol. 146, 2008
307
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2008 American Society of Plant Biologists. All rights reserved.
McGinn and Morel
quercetin and 3-MPA, the latter are not. Including
chlorophytic phytoplankton in our analysis as C3 controls clearly showed that these compounds do not
inhibit photosynthesis generally and that the desired
enzymes were indeed properly targeted by them. Moreover, DCDP and 3-MPA, which were both effective in
inhibiting photosynthesis in diatoms, are generally accepted to be specific inhibitors of key enzymes in the C4
pathway. Equally telling is the observation that chemically unrelated compounds, whose only connection is
that each inhibits a separate enzyme in a functional C4
pathway, all cripple carbon uptake in our model species. Our data fully support the idea that at least some
diatoms operate a C4-CCM. Presumably this feature
offered an ecological advantage to these organisms during the modern period of declining CO2 and may help
in nitrogen-limited regions of contemporary oceans.
MATERIALS AND METHODS
Phytoplankton Cultures
Cultures of the marine diatoms Thalassiosira pseudonana CCMP1335, Thalassiosira weissflogii CCMP1336, and Phaeodactylum tricornutum UTEX642 and
the marine chlorophytes Chlamydomonas sp. CCMP222 and Dunaliella tertiolecta
CCMP364 were grown in trace metal-buffered medium in polycarbonate
bottles under 90 mmol photons m22 s21 continuous illumination. Medium was
prepared from 0.2-mm-filtered Gulf Stream water supplemented with major
nutrients (10 mM PO432, 100 mM NO32, 100 mM SiO2), filter-sterilized vitamins,
and trace metals (80 nM zinc, 120 nM manganese, 50 nM cobalt, 20 nM copper,
and 0.5–1 mM iron) buffered with 100 mM EDTA. For gene expression experiments, T. pseudonana and P. tricornutum cells were grown under two different
CO2 regimes. To obtain T. pseudonana cells conditioned to high CO2, cells were
transferred into growth medium prebubbled with 1% CO2 gas in air until the
pH reached 7.2. As these cultures grew, the pH rose to 7.8, at which point the
cultures were harvested. For low CO2 conditioning, T. pseudonana cells were
harvested when the pH of the culture medium had reached 8.9. For P.
tricornutum, CO2 levels were controlled using filter-sterilized pH buffers. High
and low CO2-conditioned cells of P. tricornutum were obtained by including
5 mM Tris buffer at a pH of 7.5 and 8.6, respectively. For PEPC assays, low CO2
T. pseudonana cells were obtained by transferring cells into growth medium
buffered to pH 8.7 with 5 mM 4-(2-hydroxyethyl)piperazine-1-propanesulfonic
acid and growing the culture for nine generations. At the point of harvest, the
pH of these cultures had risen slightly to 8.8. Phytoplankton growth in
cultures was followed by enumerating cells with a Coulter counter (BeckmanCoulter). Specific growth rates (d21) were determined by fitting linear regression curves to the natural log of cell number versus time.
MIMS
The exchange of CO2 (m/z 5 44) and O2 (m/z 5 32) in concentrated phytoplankton suspensions was monitored continuously with MIMS. With the exception of D. tertiolecta, all cells were collected on 3-mm polycarbonate filters
(Poretics Corp.) under gentle vacuum and resuspended in 1 mL of assay buffer
(10 mM HEPES, pH 7.5; 3.5% [w/v] NaCl). Cells of D. tertiolecta were recalcitrant
to filtration, so they were collected by 5-min centrifugation (5,000g) at room
temperature and resuspended in 1 mL of assay buffer (as above). For PCKase
inhibition experiments, cells were preincubated in the dark for 5 min in the
presence of 3-MPA, a specific inhibitor of PCKase. After preincubation, the buffer
was supplemented with 400 mM NaHCO3 and the light was turned on. Light for
photosynthesis was obtained from a tungsten projector bulb and directed to the
cell chamber via a flexible fiber-optic cable at an intensity of 300 mmol photons
m22 s21 (about 33 growth light). For PEPC inhibition experiments using
flavonoid compounds, cells were prepared as above and resuspended in 1 mL
of assay buffer (10 mM HEPES, pH 7.6; 3.5% [w/v] NaCl). In these experiments,
the membrane-impermeable CA inhibitor acetazolamide (AZ) was added to a
final concentration of 20 mM after dissolution in dimethylsulfoxide (DMSO). T.
pseudonana and T. weissflogii produce an extracellular CA that equilibrates the
CO2 and HCO32 in the bulk medium (data not shown). When this CA is inhibited
by addition of AZ, the CO2 concentration in the assay medium, [CO2], can deviate
substantially from its equilibrium value, [CO2 eq], as a result of the active uptake
and fixation of CO2 and/or HCO32 by the cells. [CO2 eq] can be calculated based
on the equilibrium ratio of HCO32 to CO2 at the assay pH and from the HCO32
concentration remaining in the medium corrected for the amount of carbon fixed
in photosynthesis. The difference between [CO2] and [CO2 eq] provides a
quantitative estimate of the CO2 flux in or out of the cells: flux 5 kH ([CO2 eq] 2
[CO2]), where kH, the CO2 hydration rate constant, is 0.03 s21. The rate of CO2
release from respiration was estimated from the rate of dark O2 uptake measured
over a period of 5 min, assuming CO2:O2 stoichiometry of 1:1.
Flavonoids were prepared in equal volumes of DMSO and absolute ethanol.
For all inhibitor experiments, nonbiological variations in the 12CO2 signal were
corrected by normalizing to the inert gas argon (m/z 5 40). Calibration for CO2
was carried out by injecting known quantities of NaHCO3 and monitoring the
increase in the CO2 signal until equilibrium was reached in a medium of known
pH and temperature. Calibration for O2 was carried out by monitoring the
decrease in the O2 signal after addition of sodium dithionite to air-equilibrated
distilled water, assuming a concentration of 250 mM O2 at equilibrium with
water at 20°C. The cell suspension chamber was thermoregulated to 20°C by a
circulating water bath. The lower detection limit for CO2 (designated CO2 zero)
was determined by drop-wise addition of 10 M NaOH to distilled water. For
accurate estimates of assay cell densities, duplicate 10- to 20-mL aliquots of
concentrated cell suspension were sampled just prior to suspension in the
MIMS cuvette, diluted 500- to 1,000-fold into 3.5% NaCl and each counted in
triplicate in the Coulter counter.
RNA Isolation and cDNA Synthesis
Total RNA was isolated from diatom cells with TRIzol reagent (Invitrogen)
following the procedure of McGinn et al. (2003). The yield of total RNA was
quantified with the Ribogreen RNA quantitation kit (Molecular Probes) or by
measuring the A260. Individual mRNA transcripts were reverse transcribed to
cDNA with the SuperScript III kit (Invitrogen) from 2 mg of total RNA.
Complementary DNA templates were diluted 5-fold prior to amplification.
Gene Expression Analysis by Quantitative-PCR
Quantitative-PCR double-stranded DNA standards were generated as follows. PCR products were generated for the genes of interest by amplifying
cDNA prepared as described above in 25-mL reactions containing 1 mL of diluted
cDNA product, 1.5 mM MgCl2, 0.2 mM dNTPs, 12.5 pmol each of forward and
reverse primer, and 1 unit of Taq polymerase (Roche). The PCR cycling
conditions were as follows: 95°C for 1 min, followed by 35 cycles of 94°C for
15 s, 58°C for 20 s, and 72°C for 30 s. Amplicons were separated on 2% agarose
gels and visualized by ethidium bromide staining on a standard UV transilluminator. A single PCR product of the correct size was excised with a sterile
razor blade and purified (Qiaex II gel extraction kit; Qiagen). Purified DNAwas
quantified with a Picogreen double-stranded DNA quantitation kit (Molecular
Probes) and copy numbers per unit volume were determined. All quantitativePCR reactions were carried out on the Stratagene MX3000P thermal cycler
instrument. A SYBR Green master mix (Brilliant SYBR Green qPCR Mastermix;
Stratagene) was used for quantitative-PCR in 50-mL reactions containing 3 mL
of diluted cDNA and 25 pmol each of forward and reverse primer using the
following cycling program: 95°C for 10 min followed by 40 cycles of 94°C, 58°C
for 20 s, and 72°C for 30 s. Standards corresponding to between 102 to 106 copies
per well were amplified on the same 96-well plate as cDNA generated from
experimental material. To correct for differences in RNA starting material and
variations in cDNA synthesis efficiency (Parker and Armbrust, 2005), the
abundance of each transcript was normalized to the abundance of the actin
transcript (copies transcript of interest/copy actin), which was not biologically
regulated with changes in pCO2 tension (data not shown). Oligonucleotide
primers used for quantitative-PCR analyses are given in Table III.
PEPC Activity Assays
PEPC assays were performed according to the method of Cassar and Laws
(2007). T. pseudonana cells were collected on a 3-mm polycarbonate filter under
gentle vacuum and resuspended in 2 mL of extraction buffer containing 50 mM
BICINE (pH 7.5), 1.5 M glycerol, 10 mM MgCl2, 1 mM EDTA, 10 mM NaHCO3,
5 mM dithiothreitol, and 5 mg mL21 bovine serum albumin. Resuspended cells
were placed in a mortar chilled to 220°C and ground manually with a pestle
continuously on ice for two consecutive periods of 7 min each. After grinding,
308
Plant Physiol. Vol. 146, 2008
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2008 American Society of Plant Biologists. All rights reserved.
Carbon-Concentrating Mechanisms in Diatoms
the extract was diluted with an additional 2 mL of buffer, mixed thoroughly by
pipetting, and placed in a 30-mL glass vial and preincubated on a rotary
shaking table at room temperature for 45 min. Immediately after preincubation, three 1.2-mL aliquots of crude extract were transferred into 30-mL glass
vials and 3 mCi of NaH14CO3 was added to each. To one vial, 0.2 mM of the
PEPC inhibitor quercetin was added just prior to the addition of radiolabel.
PEP was added to a final concentration of 5 mM. Negative controls containing
no added PEP were also run. After 1 h, the reactions were terminated by
addition of 1% (v/v) 9.8 N HCl, swirled gently, and placed in a fume hood to
evaporate. Dried extracts were redissolved in 1 mL of deionized water and
counted for radioactivity after addition of 10 mL scintillation cocktail (Ultima
Gold; Packard Instrument Company).
ACKNOWLEDGMENTS
We thank Colin Jenkins (CSIRO, Canberra, Australia) for providing us
with the DCDP. We thank Dr. Nicolas Cassar (Princeton) for helpful discussions throughout this work. We also thank two anonymous reviewers for
their helpful comments and suggestions.
Received October 9, 2007; accepted November 6, 2007; published November 9,
2007.
LITERATURE CITED
Allen AE, Vardi A, Bowler C (2006) An ecological and evolutionary context
for integrated nitrogen metabolism and related signaling pathways in
marine diatoms. Curr Opin Plant Biol 9: 264–273
Badger MR, Andrews TJ, Whitney SM, Ludwig M, Yellowlees DC, Leggat
W, Price GD (1998) The diversity and coevolution of RubisCO, plastids,
pyrenoids and chloroplast based CO2 concentrating mechanisms in
algae. Can J Bot 76: 1052–1071
Badger MR, Spalding MH (2000) CO2 acquisition, concentration and
fixation in cyanobacteria and algae In RC Leegood, TD Sharkey, S von
Caemmerer, eds, Photosynthesis: Physiology and Metabolism. Kluwer
Academic Publishers, Dordrecht, The Netherlands, pp 369–397
Beardall J, Morris I (1975) Effects of environmental factors on photosynthesis patterns in Phaeodactylum tricornutum (Bacillariophyceae). II.
Effect of oxygen. J Phycol 11: 430–434
Beardall J, Mukerji D, Glover HE, Morris I (1976) The path of carbon in
photosynthesis in marine phytoplankton. J Phycol 12: 409–417
Burkhardt S, Amoroso G, Riebesell U, Sültemeyer D (2001) CO2 and
HCO32 uptake in marine diatoms acclimated to different CO 2 concentrations. Limnol Oceanogr 46: 1378–1391
Cabello-Pasini A, Swift H, Smith GJ, Alberte RS (2001) Phosphoenolpyruvate carboxykinase from the marine diatom Skeletonema costatum
and the phaeophyte Laminaria setchelli. II. Immunological characterization and subcellular localization. Bot Mar 44: 199–207
Cassar N, Laws EA (2007) Potential contribution of b-carboxylases to photosynthetic carbon isotope fractionation in a marine diatom. Phycologia
46: 307–314
Clark DR, Flynn KJ (2000) The relationship between dissolved inorganic
carbon concentration and growth rate in marine phytoplankton. Proc R
Soc Lond B Biol Sci 267: 953–959
Elrifi IR, Turpin DH (1986) Nitrate and ammonium induced suppression in
N-limited Selanastrum minutum. Plant Physiol 81: 273–279
Falkowski PG, Raven JA (1997) Aquatic Photosynthesis, Ed 1. Blackwell
Science, Malden, MA
Gibon Y, Blaesing OE, Hannemann J, Carillo P, Höhne M, Hendriks JHM,
Palacios N, Cross J, Selbig J, Stitt M (2004) A robot-based platform
to measure multiple enzyme activities in Arabidopsis using a set of cycling assays: comparison of changes in enzyme activities and transcript
levels during diurnal cycles and in prolonged darkness. Plant Cell 16:
3304–3325
Granum E, Raven JA, Leegood RC (2005) How do marine diatoms fix
10 billion tones of inorganic carbon every year? Can J Bot 83:
898–908
Guy RD, Vanlerberghe GC, Turpin DH (1989) Significance of phosphoenolpyruvate carboxylase during ammonium assimilation: carbon iso-
tope discrimination in photosynthesis and respiration by the N-limited
green alga Selenastrum minutum. Plant Physiol 89: 1150–1157
Jenkins CLD, Furbank RT, Hatch MD (1989) Inorganic carbon diffusion between C4 mesophyll and bundle sheath cells. Plant Physiol 91:
1356–1363
Jenkins CLD, Harris RLN, McFadden HG (1987) 3,3-Dichloro-2-dihydroxyphosphinoylmethyl-2-propenoate, a new, specific inhibitor of phosphoenolpyruvate carboxylase. Biochem Int 14: 219–226
Kaplan A, Reinhold L (1999) CO2-concentrating mechanisms in photosynthetic microorganisms. Annu Rev Plant Physiol Plant Mol Biol 50:
539–570
Lang DR, Racker E (1974) Effects of quercetin and F1 inhibitor on mitochondrial ATPase and energy-linked reactions in submitochondrial particles. Biochim Biophys Acta 333: 180–186
McGinn PJ, Price GD, Maleszka R, Badger MR (2003) Inorganic carbon
limitation and light control the expression of transcripts related to the
CO2-concentrating mechanism in the cyanobacterium Synechocystis sp.
strain PCC6803. Plant Physiol 132: 218–229
Moraes TF, Plaxton WC (2000) Purification and characterization of phosphoenolpyruvate carboxylase from Brassica napus (rapeseed) suspension cell cultures. Implications for phosphoenolpyruvate carboxylase
regulation during phosphate starvation and the integration of glycolysis
with nitrogen assimilation. Eur J Biochem 267: 4465–4476
Morel FMM, Cox EH, Kraepiel AML, Lane TW, Milligan AJ, Schaperdoth
I, Reinfelder JR, Tortell PD (2002) Acquisition of inorganic carbon by
the marine diatoms Thalassiosira weissflogii. Funct Plant Biol 29: 301–308
Moroney JV, Husic HD, Tolbert NE, Kitayama M, Manuel LJ, Togasaki
RK (1989) Isolation and characterization of a mutant of Chlamydomonas
reinhardtii deficient in the CO2 concentrating mechanism. Plant Physiol
89: 897–903
Pairoba CF, Columbo SL, Andreo CS (1996) Flavonoids as inhibitors of
NADP-malic enzyme and PEP carboxylase from C4 plants. Biosci Biotechnol Biochem 60: 779–783
Parker MS, Armbrust EV (2005) Synergistic effects of light, temperature,
and nitrogen source of transcription of genes for carbon and nitrogen
metabolism in the centric diatom Thalassiosira pseudonana (Bacillariophyceae).
J Phycol 41: 1142–1153
Parvathi K, Bhagwat AS, Raghavendra AS (1998) Modulation by bicarbonate of catalytic and regulatory properties of C4 phosphoenolpyruvate
carboxylase from Amaranthus hypochondriacus: desensitization to malate
and glucose 6-phosphate and sensitization to Mg(21). Plant Cell Physiol 39:
1294–1298
Rathnam CKM, Edwards GE (1977) C4-dicarboxylic acid metabolism in
bundle-sheath chloroplasts, mitochondria and strands of Eriochloa
borumensis Hack., a phosphoenolpyruvate-carboxykinase type C4 species.
Planta 133: 135–144
Ray TB, Black CC (1976) Inhibition of oxaloacetate decarboxylation during C4 photosynthesis by 3-mercaptopicolinic acid. J Biol Chem 251:
5824–5826
Ray TB, Black CC (1977) Oxaloacetate as the source of carbon dioxide for
photosynthesis in bundle sheath cells of the C4 species Panicum maximum. Plant Physiol 60: 193–196
Reinfelder JR, Kraepiel AML, Morel FMM (2000) Unicellular C4 photosynthesis in a marine diatom. Nature 407: 996–999
Reinfelder JR, Milligan AJ, Morel FMM (2004) The role of the C4 pathway
in carbon accumulation and fixation in a marine diatom. Plant Physiol
135: 2106–2111
Reiskind JB, Bowes G (1991) The role of phosphoenolpyruvate carboxykinase in a marine macroalga with C4-like photosynthetic characteristics. Proc Natl Acad Sci USA 88: 2883–2887
Roberts K, Granum E, Leegood RC, Raven JR (2007) C3 and C4 pathways of
photosynthetic carbon assimilation in marine diatoms are under genetic, not environmental, control. Plant Physiol 145: 230–235
Rost B, Riebesell U, Burkhardt S, Sültemeyer D (2003) Carbon acquisition
of bloom-forming marine phytoplankton. Limnol Oceanogr 48: 55–67
Tortell PD (2000) Evolutionary and ecological perspectives on carbon
acquisition in phytoplankton. Limnol Oceanogr 45: 744–750
Tortell PD, Martin CL, Corkum ME (2006) Inorganic carbon uptake and
intracellular assimilation by subarctic Pacific phytoplankton assemblages. Limnol Oceanogr 51: 2102–2110
Tortell PD, Reinfelder JR, Morel FMM (1997) Active uptake of bicarbonate
by diatoms. Nature 390: 243–244
Plant Physiol. Vol. 146, 2008
309
Downloaded from on June 17, 2017 - Published by www.plantphysiol.org
Copyright © 2008 American Society of Plant Biologists. All rights reserved.