carbon acquisition by diatoms

10
REVIEW Carbon acquisition by diatoms Karen Roberts Espen Granum Richard C. Leegood John A. Raven Received: 31 August 2006 / Accepted: 11 April 2007 / Published online: 12 May 2007 Ó Springer Science+Business Media B.V. 2007 Abstract Diatoms are responsible for up to 40% of pri- mary productivity in the ocean, and complete genome sequences are available for two species. However, there are very significant gaps in our understanding of how diatoms take up and assimilate inorganic C. Diatom plastids origi- nate from secondary endosymbiosis with a red alga and their Form ID Rubisco (ribulose-1,5-bisphosphate carbox- ylase-oxygenase) from horizontal gene transfer, which means that embryophyte paradigms can only give general guidance as to their C acquisition mechanisms. Although diatom Rubiscos have relatively high CO 2 affinity and CO 2 /O 2 selectivity, the low diffusion coefficient for CO 2 in water has the potential to restrict the rate of photosynthesis. Diatoms growing in their natural aquatic habitats operate inorganic C concentrating mechanisms (CCMs), which provide a steady-state CO 2 concentration around Rubisco higher than that in the medium. How these CCMs work is still a matter of debate. However, it is known that both CO 2 and HCO 3 are taken up, and an obvious but as yet unproven possibility is that active transport of these species across the plasmalemma and/or the four-membrane plastid enve- lope is the basis of the CCM. In one marine diatom there is evidence of C 4 -like biochemistry which could act as, or be part of, a CCM. Alternative mechanisms which have not been eliminated include the production of CO 2 from HCO 3 at low pH maintained by a H + pump, in a compartment close to that containing Rubisco. Keywords Active transport Á Bacillariophyceae Á Carbon dioxide Á CO 2 -concentrating mechanism Á Carbonic anhydrase Á C 4 photosynthesis Á Form ID Rubisco Á Pyrenoid Introduction There are at least 10,000 extant species of diatoms, con- stituting the class Bacillariophyceae (phylum Hetero- kontophyta, kingdom Chromista: Falkowski et al. 2004). Diatoms occur in the benthos and plankton of the ocean and inland waters, and account for up to 40% of the total marine primary productivity of approximately 50 Pg C per year (Field et al. 1998; Granum et al. 2005). Complete genome sequences for the marine diatoms Thalassiosira pseudonana and Phaeodactylum tricornutum have revealed a unique mixture of characteristics, as befits organisms derived from secondary endosymbiosis, in a clade separate from plants and from opisthokonts (animals and fungi) (Baldauf 2003; Armbrust et al. 2004; Montsant et al. 2005; Allen et al. 2006). A unique characteristic of diatoms is the bipartite, silicified cell wall which greatly influences their ecology (Raven and Waite 2004). In chromists, the endosymbiotic origin of plastids from red algal cells (Falkowski et al. 2004) accounts for their two additional membranes (‘plastid endoplasmic reticu- lum’). These additional membranes require different, incompletely characterised plastid targeting sequences from those found in green and red algae and embryophytes with just the two ‘normal’ plastid envelope membranes, making it difficult to predict if a given nuclear-encoded protein is targeted to the plastids (Apt et al. 2002). The additional membranes also have implications for the K. Roberts Á J. A. Raven (&) Plant Research Unit, University of Dundee at SCRI, Scottish Crop Research Institute, Invergowrie, Dundee DD2 5DA, UK e-mail: [email protected] E. Granum Á R. C. Leegood Department of Animal and Plant Sciences, University of Sheffield, Sheffield S10 2TN, UK 123 Photosynth Res (2007) 93:79–88 DOI 10.1007/s11120-007-9172-2

Upload: karen-roberts

Post on 15-Jul-2016

215 views

Category:

Documents


3 download

TRANSCRIPT

REVIEW

Carbon acquisition by diatoms

Karen Roberts Æ Espen Granum Æ Richard C. Leegood ÆJohn A. Raven

Received: 31 August 2006 / Accepted: 11 April 2007 / Published online: 12 May 2007

� Springer Science+Business Media B.V. 2007

Abstract Diatoms are responsible for up to 40% of pri-

mary productivity in the ocean, and complete genome

sequences are available for two species. However, there are

very significant gaps in our understanding of how diatoms

take up and assimilate inorganic C. Diatom plastids origi-

nate from secondary endosymbiosis with a red alga and

their Form ID Rubisco (ribulose-1,5-bisphosphate carbox-

ylase-oxygenase) from horizontal gene transfer, which

means that embryophyte paradigms can only give general

guidance as to their C acquisition mechanisms. Although

diatom Rubiscos have relatively high CO2 affinity and

CO2/O2 selectivity, the low diffusion coefficient for CO2 in

water has the potential to restrict the rate of photosynthesis.

Diatoms growing in their natural aquatic habitats operate

inorganic C concentrating mechanisms (CCMs), which

provide a steady-state CO2 concentration around Rubisco

higher than that in the medium. How these CCMs work is

still a matter of debate. However, it is known that both CO2

and HCO3– are taken up, and an obvious but as yet unproven

possibility is that active transport of these species across

the plasmalemma and/or the four-membrane plastid enve-

lope is the basis of the CCM. In one marine diatom there is

evidence of C4-like biochemistry which could act as, or be

part of, a CCM. Alternative mechanisms which have not

been eliminated include the production of CO2 from HCO3–

at low pH maintained by a H+ pump, in a compartment

close to that containing Rubisco.

Keywords Active transport � Bacillariophyceae �Carbon dioxide � CO2-concentrating mechanism �Carbonic anhydrase � C4 photosynthesis � Form ID

Rubisco � Pyrenoid

Introduction

There are at least 10,000 extant species of diatoms, con-

stituting the class Bacillariophyceae (phylum Hetero-

kontophyta, kingdom Chromista: Falkowski et al. 2004).

Diatoms occur in the benthos and plankton of the ocean

and inland waters, and account for up to 40% of the total

marine primary productivity of approximately 50 Pg C per

year (Field et al. 1998; Granum et al. 2005). Complete

genome sequences for the marine diatoms Thalassiosira

pseudonana and Phaeodactylum tricornutum have revealed

a unique mixture of characteristics, as befits organisms

derived from secondary endosymbiosis, in a clade separate

from plants and from opisthokonts (animals and fungi)

(Baldauf 2003; Armbrust et al. 2004; Montsant et al. 2005;

Allen et al. 2006). A unique characteristic of diatoms is the

bipartite, silicified cell wall which greatly influences their

ecology (Raven and Waite 2004).

In chromists, the endosymbiotic origin of plastids from

red algal cells (Falkowski et al. 2004) accounts for their

two additional membranes (‘plastid endoplasmic reticu-

lum’). These additional membranes require different,

incompletely characterised plastid targeting sequences

from those found in green and red algae and embryophytes

with just the two ‘normal’ plastid envelope membranes,

making it difficult to predict if a given nuclear-encoded

protein is targeted to the plastids (Apt et al. 2002). The

additional membranes also have implications for the

K. Roberts � J. A. Raven (&)

Plant Research Unit, University of Dundee at SCRI, Scottish

Crop Research Institute, Invergowrie, Dundee DD2 5DA, UK

e-mail: [email protected]

E. Granum � R. C. Leegood

Department of Animal and Plant Sciences, University of

Sheffield, Sheffield S10 2TN, UK

123

Photosynth Res (2007) 93:79–88

DOI 10.1007/s11120-007-9172-2

transfer of inorganic C species, and other low Mr solutes,

into plastids. Replacement of the Form IB Rubisco inher-

ited from the red algal plastid ancestor by a Form ID Ru-

bisco in chromists probably occurred by horizontal gene

transfer not directly related to secondary endosymbiosis.

With this background we consider the implications of

the available data for inorganic C acquisition by diatoms

and early stages of organic C metabolism (Raven et al.

2007).

Diatom Rubisco and the need for a CCM

Form ID Rubiscos have relatively high affinities for CO2,

high CO2/O2 selectivities, and a relatively low CO2-satu-

rated specific reaction rate (Badger et al. 1998; Tcherkez

et al. 2006). While diatom selectivity and affinity values

are not the highest measured for Form ID Rubiscos, they

are higher than those in coccolithophores (Badger et al.

1998; Whitney et al. 2001; Shiraiwa et al. 2004). Form ID

Rubiscos generally have higher CO2/O2 selectivities than

the other Rubiscos found in algal cells.

It has long been known that diatom cells have a very

high affinity for inorganic C (Barker 1935). While Barker

(1935) used Warburg buffers with very high inorganic C

concentrations and high, and varying, pH values to obtain

the required CO2 concentrations, such conditions could be

experienced in nature by diatoms in saline, high-carbonate,

inland waters. Badger et al. (1998) (see also Burkhardt

et al. 2001; Rost et al. 2003) showed that the affinity for

inorganic C expressed as CO2 in photosynthesis by diatom

cells can be forty times that of isolated diatom Rubisco. To

explain this with diffusive CO2 entry there would have to

be at least a 40-fold excess of inorganic C-saturated Ru-

bisco activity over the inorganic C-saturated rate of pho-

tosynthesis in vivo, both expressed on a cell basis (Raven

1984; Badger et al. 1998). In this case inorganic C-satu-

rated photosynthesis in the cells would be limited by

reactions such as reduction of NADP+ and phosphorylation

of ADP (Raven 1984). ‘At least’ refers to the 10,000-fold

lower diffusion coefficient of CO2 in water compared to

that in air, and in even the smallest diatoms the restriction

imposed by aqueous phase diffusion exceeds that in C3

land plants (Raven 1984). Such a 40-fold, or more, excess

of Rubisco protein and activity is practically impossible.

Assuming only a two-fold excess of Rubisco activity over

that required for the observed inorganic C-saturated rate of

photosynthesis in vivo, to half saturate Rubisco would re-

quire a two-fold accumulation of CO2 at the Rubisco active

site relative to that in normal seawater (~2 mol m–3 inor-

ganic C) at pH 8.2. A 20-fold CO2 accumulation would be

needed with 1/10 the normal inorganic C concentration in

seawater, a value in the range used in many experiments on

CO2 accumulation.

The observed ratio of internal to external inorganic C

concentration in marine diatom cells during photosynthesis

is up to 3.5 in a medium with more than twice the seawater

inorganic C concentration at pH 7.5 (Colman and Rotatore

1995), and is higher than 3.5 at lower external concentra-

tions at pH 7.5–8.0 (Burns and Beardall 1987; Rotatore

et al. 1995; Johnston and Raven 1996; Mitchell and

Beardall 1976; Reinfelder et al. 2004). Even higher ratios

occur in freshwater diatoms (Rotatore and Colman 1992).

Relating these findings to the CO2 concentration at the

active site of Rubisco involves assumptions about the

intracellular location of the accumulated inorganic C pool

and the pH at this site. Table 1 shows calculations

assuming that there is an equilibration of CO2 across the

plasmalemma, but not across the plastid envelope, and vice

versa. The calculations are for two external inorganic

carbon concentrations and for a steady-state CO2 concen-

tration in the plastid stroma that is twice the concentration

in normal seawater. For any of the assumptions made the

predicted inorganic C concentration averaged over the

whole cell volume (Table 1) is consistent with measured

mean intracellular accumulations. Even a requirement for a

CO2

concentration in the plastid stroma which is two times

that found in normal seawater would still yield predicted

inorganic C concentrations averaged over the whole cell

that are within the range of measured values. Hence, sig-

nificant inorganic C accumulation inside the chloroplast,

and even the cytoplasm, can be obtained without a large

increase in mean intracellular concentration due to the high

volume fraction of vacuole with little inorganic C. How-

ever, some of the diatoms that are widely used in experi-

ments (e.g., P. tricornutum) probably have a lower fraction

of the protoplast occupied by vacuole than the organisms

considered in Table 1, and so would have higher predicted

mean intracellular inorganic C concentrations. The range

of cells considered in Table 1 was restricted by the avail-

ability of morphometric data. We now consider the pro-

posed diatom CCMs which could deliver the required

increase in CO2 concentration around Rubisco.

Proposals for diatom CCMs

Table 2 and Fig.1 give an overview of possible diatom

CCMs, based on the CCMs already known, or proposed,

for O2-producing photosynthetic organisms. Mechanisms

Ia-c in Table 2 are based on active transport of CO2, HCO3–

and H+ across membranes, i.e., vector energized reactions.

Mechanisms IIa-c in Table 2 are based on energized

reactions in aqueous solution, i.e., scalar energized reac-

80 Photosynth Res (2007) 93:79–88

123

tions, although IIb also requires active transport for the

night-time accumulation of malic acid in vacuoles.

Active influx of CO2 (Table 2, Ia) has only been found

in eukaryotes, where it occurs at the plasmalemma and/or

the plastid envelope (Giordano et al. 2005). Diatoms can

take up CO2 from the medium (reviewed by Giordano et al.

2005) with CO2 accumulation in the cells (Rotatore and

Colman 1992). However, it is not known where the active

transport of CO2 occurs, i.e., across the plasmalemma or

the plastid envelope (see Raven 1997, and Table 2, Icii).

Wittpoth et al. (1998) showed that isolated plastids from

two species of diatom had over a third (on a chlorophyll

basis) of the intact cell rates of inorganic C-dependent

photosynthetic O2 evolution, but did not determine the

inorganic C species taken up, or whether inorganic C was

accumulated by the plastids. Since almost all the isolated

plastids had lost their chloroplast endoplasmic reticulum

membranes, their in vivo relevance is unclear. Uptake of

HCO3– rather than CO2 into the chloroplasts would facilitate

the operation of the mechanism of pyrenoid function de-

scribed below involving conversion of HCO3– from the

stroma to CO2 in the pyrenoid.

Active influx of HCO3– (Table 2, Ib) occurs, using sev-

eral mechanisms, at the plasmalemma of cyanobacteria.

HCO3– transport also occurs at the plasmalemma and plastid

envelope of eukaryotes, and leads to accumulation at both

the plastid and the whole cell level. HCO3– is taken up by

almost (see Cassar et al. 2002) all diatoms (Giordano et al.

2005), but it is not known whether it is transported across

the plasmalemma or plastid envelope. Despite the inside-

negative electrical potential (Boyd and Gradmann 1999)

across the plasmalemma of diatoms, 1:1 symport with H+

or 1:1 antiport with OH– would permit HCO3– entry without

energy input (see Raven 1997, and Table 2, Icii). Operation

of the mechanism of pyrenoid function discussed below

would be facilitated by HCO3– rather than CO2 influx from

the cytosol into the plasid stroma.

The required CO2 concentration around Rubisco is

more readily accommodated if the observed inorganic C

accumulation averaged over the whole cell volume is

recalculated for accumulation only in the plastids. Hence,

studies of inorganic C transport across the plastid enve-

lope of diatoms, and further investigation of mechanisms

based on active H+ transport (Raven 1997) are urgently

needed.

If there is accumulation of HCO3– in the plastid stroma

then a mechanism of CO2 supply to Rubisco of the kind

believed to occur in cyanobacteria could be envisaged,

with the pyrenoids found in the great majority of diatoms

(Schmid 2001) taking the place of carboxysomes. A diatom

version may have active HCO3– influx across the plastid

envelope and/or the plasmalemma, yielding a higher con-

centration of HCO3– in the chloroplast stroma than is found

in the medium, with Rubisco and carbonic anhydrase (CA)

mainly localized in the pyrenoid. In this model, HCO3–

enters the pyrenoid with ribulose bisphosphate (RuBP),

CO2 generated by CA is assimilated by Rubisco, and the

resulting phosphoglycerate (PGA) diffuses into the stroma

where it is further metabolised by the photosynthetic C

reduction cycle (PCRC). Diatom pyrenoids are typically

surrounded by a membrane of unknown composition

(Schmid 2001); is this membrane analogous to the pro-

teinaceous shell of carboxysomes (Kerfeld et al. 2005),

with anion-selective pores permitting fluxes of the sub-

strates HCO3– and RuBP and the product PGA? Kerfeld

et al. (2005) suggests that these channels restrict loss of the

CO2 generated by CA and entry of O2 that competes with

CO2 at the active site of Rubisco.

Table 1 Model calculations of total inorganic C accumulation in four

diatoms based on non-energized (= passive) CO2 equilibration across

the plasmalemma and tonoplast with active transport at the chloro-

plast envelope, or CO2 equilibration across the chloroplast envelope

and tonoplast with active transport at the plasmalemma

Organism Diatomatenue

Cyclotellameneghiniana

Fragilariacapucina

Stephanodiscusbinderanus

Protoplast volume 230 lm3 250 lm3 320 lm3 760 lm3

Vacuole volume as fraction of protoplast volume 0.45 0.40 0.45 0.62

Cytoplasm volume as fraction of protoplast volume 0.35 0.37 0.32 0.24

Chloroplast volume as fraction of protoplast volume 0.20 0.23 0.23 0.14

Mean [inorganic C] in protoplast for active transport at the chloroplast envelope and

external [inorganic C] of (i) 0.2 mol m–3 or (ii) 2 mol m–3(i) 0.82 (i) 0.94 (i) 0.94 (i) 0.57

(ii) 0.97 (ii) 1.10 (ii) 1.07 (ii) 0.68

Mean [inorganic C] in protoplast for active transport at the plasmalemma and external

[inorganic C] of 0.2 or 2 mol m–31.13 1.27 1.23 0.80

Other assumptions made were chloroplast stromal [inorganic C] of 4 mol m–3 and external medium [inorganic C] of either 0.2 or 2 mol m–3.

Quantitative morphometry yielding compartmental volume fractions was obtained from Sicko-Goad et al. (1977, 1988) and Sicko-Goad and

Stoermer (1979). Compartmental pH values (seawater medium 8.2; vacuole 5.0; chloroplast stroma 8.2; cytoplasm 7.6) were obtained from

Raven (1984), Johnson and Raven (1996) and Nimer et al. (1999, for a marine dinoflagellate). At equilibrium the ratios of [inorganic C] between

compartments of pH 8.2, 7.6 and 5.0 are 1:0.23:0.0065

Photosynth Res (2007) 93:79–88 81

123

It must be admitted that there is no evidence either for

the occurrence of CA in diatom pyrenoids, or of the pro-

posed properties of the pyrenoid membrane. However,

there is good evidence for the occurrence of a b-CA in

aggregations in a ring at the periphery (near the girdle

lamellae) of the stroma in a strain of P. tricornutum which

lacks pyrenoids (Tanaka et al. 2005). These b-CA aggre-

gates seem to be involved in the CCM (Satoh et al. 2001);

do they contain Rubisco and function as membrane-less

pyrenoids (Tanaka et al. 2005)? Less is known about the

Table 2 Summary of proposed CCMs, with their relevance to diatoms

Mechanism Location in cell Phylogenetic distribution Occurrence in

diatoms?

References

Ia Active influx of CO2 Plasmalemma and/or plastid

envelope; HCO3– rather than

CO2 influx from the cytosol

into the plastid stroma

would be more compatible

with present views of

pyrenoid function

Eukaryotic algae (but see Icii).

Hornworts? Some aquatic

flowering plants?

Yes, provided

mechanisms Icii

and Iciii are not

operative, but not

clear where active

transport occurs

Colman et al. (2002),

Burkhardt et al. (2001),

Cassar et al. (2002), Rost

et al. (2003)

Ib Active influx of

HCO3–

Plasmalemma and/or plastid

envelope; HCO3– rather than

CO2 influx from the cytosol

into the plastid stroma

would be more compatible

with present views of

pyrenoid function

Cyanobacteria. Eukaryotic

algae (but see Icii).

Hornworts? Some aquatic

flowering plants?

Yes, provided

mechanisms Icii

and Iciii are not

operative, but not

clear where active

transport occurs

Badger et al. (2006), Colman

et al. (2002), Burkhardt

et al. (2001), Rost et al.

(2003) cf. Laws and Cassar

(2002)

Ici Active efflux of H+

to cell wall,

generation of CO2

from HCO3–

Plasmalemma; CO2 generated

in cell wall, diffuses into

plastid stroma

Some characeans and some

freshwater aquatic plants;

perhaps some marine

macrophytes

Unlikely; only

possible in larger

cells

Walker et al. (1980)

Icii Active influx of H+

to thylakoid lumen,

generation of CO2

from HCO3–

Thylakoid membrane; HCO3–

from stroma enters

thylakoid lumen where CO2

is generated, and then

diffuses into the stroma and,

specifically, the pyrenoid

Chlamydomonas reinhardtii? Possibly; not readily

compatible with

mechanisms Ia or

Ib

Pronina and Semenenko

(1992), Raven (1997),

Hanson et al. (2003), Mitra

et al. (2005)

Iciii Active efflux of H+

from cytosol or

stroma to plastid

inter-membrane

space, generation of

CO2 from HCO3–

Membranes of chloroplast

endoplasmic reticulum; CO2

generated in inter-

membrane space, diffuses

into stroma

Some eukaryotic algae with

plastids obtained by

secondary endo-symbiosis?

Possibly; not readily

compatible with

mechanisms Ia or

Ib

Lee and Kugrens (1998,

2000)

IIa Single-cell C4

biochemistry

(C3 + C1) carboxylation of a

C3 carboxylic acid occurs in

the cytosol and (C4 – C1)

decarboxylation of a C4 acid

occurs in the plastids, with

refixation of CO2 by

Rubisco

Single-cell C4 in some

terrestrial and aquatic

flowering plants, and

marine algae such as.

Udotea flabellum and

(perhaps) Emiliania huxleyi

Thalassiosiraweissflogiisuggested to be C4;

more likely C3 –

C4 intermediate

Reinfelder et al. (2000, 2004),

Johnston et al. (2001),

Morel et al. (2002),

Edwards et al. (2004),

Granum et al. (2005), Tsuji

et al. (2006)

IIb Crassulacean acid

metabolism

(C3 + C1) carboxylation occurs

in the cytosol at night, malic

acid is stored in the vacuole,

and (C4 – C1)

decarboxylation occurs in

the cytosol during day, with

refixation of CO2 by

Rubisco

Terrestrial and aquatic ferns,

lycopods and flowering

plants

Possibly, but only in

larger cells with a

high vacuole:

cytoplasm ratio

Keeley and Rundel (2003),

Holtum et al. (2005)

IIc Passive CO2 entry,

active conversion to

HCO3– on thylakoid

membrane

Active generation of HCO3– in

cytosol, transported into the

carboxysomes

Cyanobacteria Possibly (on

chloroplast

endoplasmic

reticulum)

Tchernov et al. (2001),

Badger et al. (2006)

See also Giordano et al. (2005) and Raven et al. (2007)

82 Photosynth Res (2007) 93:79–88

123

role of the other CA found in diatoms, including two other

novel CAs, one able to use either Zn or Co as cofactor, the

other specific for Cd (Roberts et al. 1997; Cox et al. 2000;

Lane et al. 2005; Park et al. 2007; Szabo and Colman

2007).

The final category of CCMs based on active transport

(Table 2, Ic) relies on acidification of a compartment

accessible to HCO3– from the medium; formation of the

equilibrium (high) CO2: HCO3– ratio, usually requiring CA

activity, is followed by diffusion of CO2 to the compart-

ment containing Rubisco. A mechanism (Table 2, Ici)

where such acidification occurs in zones at the surface of

diatom cells is unlikely in view of their relatively small

size (Raven 1984), and the limited expression of extra-

cellular CA in diatoms in nature (Martin and Tortell 2006).

Intracellular variants on this acidification theme include

CO2 generation in the thylakoid lumen (Table 2, Icii). Here

the suggestion is that HCO3–, ultimately from the medium,

enters the thylakoid lumen of the plastid through anion

channels, and a CA then generates CO2 which diffuses to

Rubisco in the stroma, or, more specifically, the pyrenoid

(Pronina and Semenenko 1992; Raven 1997). In agreement

with this hypothesis the freshwater green microalga Chla-

mydomonas reinhardtii has a luminal a-CA (Cah3) essen-

tial for the CCM located in thylakoid tubules in the

pyrenoid (Hanson et al. 2003; Mitra et al. 2005). However,

the lack of any detectable HCO3– channels in the thylakoids

of three strains of green algae with CCMs, including a

wall-less strain of C. reinhardtii (van Hunnik and Sulte-

meyer 2002), is clearly a problem. Could this mechanism

occur in diatoms? No Cah3 analogue has been found in

diatoms, but the ubiquitous luminal photosystem II protein

PsbO has CA activity (Enami et al. 2005; Lu et al. 2005;

Hillier et al. 2006; cf. Stemler and Govindjee 1973).

Nevertheless, the CCM of C. reinhardtii is absolutely

dependent upon the expression of Cah3 (Hanson et al.

2003). Furthermore, a few diatoms have no pyrenoids, and,

when pyrenoids do occur, they do not always have thy-

lakoids running through them. However, the ‘mem-

brane’(when present) around the pyrenoids could represent

the ends of thylakoids (Drum 1963; Drum and Pancratz

1964; Schmid 2001; Mayama et al. 2004).

The hypothesis of Lee and Kugrens (1998, 2000) (Ta-

ble 2, Iciii), with the acidified compartment that generates

CO2 identified as the chloroplast endoplasmic reticulum,

has not been explicitly tested on diatoms, or any of the

other algae to which it may apply. The hypothesis could be

tested when photosynthetically active diatom plastids can

be isolated with their chloroplast endoplasmic reticulum

intact (see Wittpoth et al. 1998).

The possibility that diatoms have C4-like photosynthesis

(Table 2, IIa) was first suggested in the 1970s when short-

term (2s) inorganic 14C supply experiments (Beardall et al.

1976) showed labelling of malate, typical of C4 photo-

synthesis, rather than PGA and sugar phosphates, typical of

C3 biochemistry. Furthermore, assays of the in vitro

activity of carboxylases showed a relatively high ratio of

phosphoenolpyruvate carboxylase (PEPC) to Rubisco.

Other, usually rather longer-term, labelling experiments

showed a predominance of PGA and sugar phosphates

(Coombs and Volcani 1968; Holdsworth and Colbeck

1976; Kremer and Berks 1978; Mortain-Bertrand et al.

1987). The implicit or explicit assumption that diatoms

have only C3 photosynthetic biochemistry went unchal-

lenged until Reinfelder and colleagues (Reinfelder et al.

2000, 2004; Morel et al. 2002; cf. Johnston et al. 2001;

Granum et al. 2005) found significant evidence of C4

photosynthesis in the marine diatom T. weissflogii.

tsalporolhcmsalpotycmsalpirep

AC AC

CPEP

OC 2OC 2

OCH 3-OCH 3

-

uR ocsib

aIOC 2

OCH 3-

eloucav

C4 dicadiokalyht

cimsalpodneuluciter m

CRCP

AC AC

C3 dica

C4 dica

aI aI

bI bI bI

cI icII

baII aII

bII

cI iii cI ii

H+

bII

AC

C4 dica

C3 dica

OC 2

OCH 3-

OC 2

OCH 3-

uR PB AGP×2

EM/KCPEP

HC[ 2 ]O

H+ H+H+

dioneryp

OCH 3-

OC 2

Fig. 1 Schematic presentation

of the possible CCMs in

diatoms (Table 2, and text). The

simultaneous occurrence of all

of these mechanisms would lead

to futile cycles; only restricted

subsets of these processes can

function together as parts of

CCMs

Photosynth Res (2007) 93:79–88 83

123

The inorganic 14C labelling results for T. weissflogii

were confirmed and extended by Roberts et al. (unpub-

lished). Their 2s and 5s labelling experiments showed a

rather lower and variable ratio of malate to PGA and sugar

phosphates than that found by Reinfelder et al. (2000), but

still with a predominance of malate. The labelling pattern

was not influenced by the growth inorganic C concentra-

tion. Overall, the labelling pattern resembled that of certain

C3 – C4 intermediate flowering plants such as Flaveria

linearis, with PGA and malate in parallel as the initial

fixation products, and subsequent label transfer from C4

acids to PGA and sugar phosphates (Monson et al. 1986).

By contrast, T. pseudonana strictly showed a C3 labelling

pattern, again regardless of the inorganic C concentration

used for growth (Roberts et al. unpublished), and little or

no up-regulation of C4-related genes by low inorganic C

concentration (Granum et al. 2005; Granum et al. unpub-

lished). Armbrust et al. (2004), in analysing the complete

genome sequence of T. pseudonana, suggested that this

diatom (like T. weissflogii) had C4 photosynthetic metab-

olism. This suggestion was based on the presence of genes

for PEPC, phosphoenolpyruvate carboxykinase (PEPCK)

and pyruvate-orthophosphate dikinase which, with plastid

envelope transporters, could allow single-celled C4 photo-

synthetic metabolism; these genes also occur in P. tricor-

nutum (Montsant et al. 2005). The metabolic labelling

experiments show that the presence of these genes, as in C3

flowering plants, is inadequate to demonstrate C4 photo-

synthetic metabolism; this is also the case for a similar

suggestion made for the picoplanktonic prasinophyte Ost-

reococcus tauri (Derelle et al. 2006). More work is clearly

needed (see Granum et al. 2005) to determine more pre-

cisely the pathways of inorganic C assimilation, and their

relative contributions to overall photosynthesis, in T.

weissflogii.

Is it unexpected to find both C3 and C3 – C4 interme-

diate photosynthetic pathways in a single morphologically

defined genus of diatoms? It is instructive to look at the

distribution of C3, C4 and C3 – C4 intermediate photo-

synthetic pathways in flowering plants. Molecular phylo-

genetic analyses show that there have been at least 31

independent origins of C4 photosynthesis among flower-

ing plants (Kellog 1999), yielding some 8,000 C4 species

(Sage et al. 1999). Sage et al. (1999) list five morpho-

logically defined genera of the Poaceae with both C3 and

C4 species, and Kellog (1999) points out that two of these

genera also have C3 – C4 intermediate species. Kellog

(1999) cautions that there is no adequate phylogeny of the

subfamily Panicoideae and that, while the genus Panicum

has C3 and C3 – C4 intermediate species as well as spe-

cies with all three subtypes of C4, Panicum is probably

paraphyletic. Interestingly, molecular phylogenetic studies

of the diatom order Thalassiosirales show that T.

pseudonana and T. weissflogii are not closely related

within the paraphyletic genus Thalassiosira (Kaczmarska

et al. 2006). Returning to the question at the beginning of

the paragraph, it seems no more unexpected to find a

diversity of photosynthetic pathways of inorganic C

assimilation in a genus of diatoms than in a genus of

flowering plants. Additional evidence of photosynthetic

variation within the genus Thalassiosira comes from work

on T. oceanica from a low-iron area in the North Pacific.

T. oceanica, unlike T. pseudonana, T. weissflogii and all

other chromists so far examined, uses plastocyanin rather

than cytochrome c6 in electron transport from the cyto-

chrome b6-f complex to P700 (Strzepek and Harrison

2004; Peers and Price 2006).

A temporally rather than spatially separated variant of

C4 photosynthesis is Crassulacean acid metabolism (CAM;

Table 2, IIb). The volume fraction of vacuoles in larger

diatom cells is similar to that of CAM vascular plants, so

there is adequate malic acid storage capacity (Raven 1987,

1995), as shown by the following calculation based on data

for the giant diatom Ethmodiscus (Villareal et al. 1999).

The organic C content is 0.48 · 10–6 mol cell–1, and the

cell volume is 2.2 · 10–9 m3. Assuming that the vacuole

contributes 90% of the cell volume (Raven 1995), and that

CAM involves a relatively modest (Raven and Spicer

1996) diel variation of vacuolar malic acid of 200 mol m–3,

the CO2 stored in malic acid overnight for refixation in the

day is 200 · 0.9 · 2.2 · 10–9 or 0.40 · 10–6 mol CO2 per

diel cycle. This is not quite adequate for one cell doubling

per day based on CAM alone, especially if CO2 losses in

leakage and respiration and dissolved organic C losses are

taken into account. However, the estimated doubling time

of Ethmodiscus cells in the North Pacific Gyre is not less

than 45 h, so CAM could be a possible means of supplying

all of the inorganic C used for growth. Nevertheless, there

is the important question of what could be the selective

advantage of CAM in the oligotrophic ocean. For the

smaller-celled diatoms that have been examined the mea-

sured rates of dark inorganic 14C fixation (e.g., Mortain-

Bertrand et al. 1998) match the anaplerotic requirement for

inorganic N assimilation (Granum and Myklestad 1999),

thus not leaving significant surplus for CAM. Furthermore,

small-celled diatoms do not have a large enough fraction of

their cell volume occupied by vacuoles (Raven 1987, 1995)

to permit CAM to support their high growth rates (Banse

1982).

The final possible CCM in diatoms involves passive

CO2 entry and energized conversion to HCO3–, a mecha-

nism only found so far in cyanobacteria (Table 2, IIc).

Plastid NADH dehydrogenases are not known in diatoms

(Peltier and Cournac 2002; Armbrust et al. 2004), so the

mechanism suggested for cyanobacteria with energized

CO2 conversion to HCO3– in the cytosol, involving a H+

84 Photosynth Res (2007) 93:79–88

123

gradient produced by a NADH dehydrogenase, cannot

occur in diatoms (Tchernov et al. 2001; Badger et al.

2006).

In conclusion, there are many missing links in the story

of diatom CCMs. None of the proposed mechanisms are

completely ‘joined up’ and, while some of them could be

complementary, e.g., active influx of CO2 (Ia) with a

cytosol CA producing HCO3– that then enters plastids (IIc),

or active influx of HCO3– (Ib) with HCO3

– using C4

metabolism (or, more likely, C3 – C4 intermediate metab-

olism) (IIa), some other combinations could result in futile

cycles. An example of such a combination is that of a

mechanism that delivers CO2 to Rubisco in the pyrenoid by

C4 photosynthesis involving PEPCK (IIa), or by CO2 dif-

fusion across the thylakoid membrane after CA-catalysed

conversion of HCO3– to CO2 in the acidic thylakoid lumen

(Icii), with mechanism Ib involving CA-catalysed conver-

sion of HCO3– to CO2 in the pyrenoid. A more general

aspect of futile cycling is that of leakage of accumulated

inorganic C, especially CO2, during the operation of

CCMs. Leakage of inorganic C from diatom cells (or

chloroplasts) is probably small as suggested by the high

photon yields (on an absorbed photon basis) of photosyn-

thesis by T. pseudonana (Myers 1980) and P. tricornutum

(Geider et al. 1985, 1986), at least in the latter case for cells

grown in air-equilibrated artificial seawater medium.

Leakage is further discussed by Granum et al. (2005) and

Raven et al. (2007)

Regulation of diatom CCMs

The affinity of diatom photosynthesis for inorganic C in-

creases with decreasing inorganic C supply (Johnston and

Raven 1996; Fielding et al. 1998; Burkhardt et al. 2001;

Rost et al. 2003). Matsuda et al. (2001, 2002) showed that

CO2 rather than HCO3– is the external inorganic C species

involved in this acclimation in P. tricornutum, as in

freshwater green microalgae, but different from that in

cyanobacteria (Badger et al. 2006). Cyclic AMP is a

mediator in the response of the CCM of P. tricornutum to

external CO2 (Harada et al. 2006).

Photorespiration in diatoms

The CCMs of diatoms do not completely suppress the

oxygenase activity of Rubisco, especially when there is

significant CO2 depletion and/or O2 accumulation relative

to air-equilibrium conditions, and/or high irradiances and

temperatures (Birmingham and Colman 1979; Colman and

Hosein 1980; Birmingham et al. 1982; Parker et al. 2004;

Parker and Armbrust 2005). Roberts et al. (unpublished)

found significant labelling of phosphoglycolate, glycolate

and glycine among the 2–5s products of inorganic 14C

assimilation by T. pseudonana and T. weissflogii under

near-ambient conditions.

Parker et al. (2004) and Parker and Armbrust (2005)

have shown increased transcription of glycine decarbox-

ylase, an enzyme of the photorespiratory C oxidation cycle

(PCOC) of embryophytes and many green algae, under

conditions which permit more Rubisco oxygenase activity

in T. pseudonana and T. weissflogii, and Granum et al.

(unpublished) have confirmed this for T. pseudonana under

low CO2 conditions. Armbrust et al. (2004) suggested that

the PCOC is the pathway of phosphoglycolate metabolism

in T. pseudonana; while there can still be doubts as to the

occurrence of all of the genes required for this cycle in the

genome, it is salutory that a bacterial like glycerate kinase

was only recently recognised as a cycle enzyme in Ara-

bidopsis thaliana (Boldt et al. 2005). It has also not proved

possible to establish at the genetic level the occurrence of

the alternative tartronic semialdehyde pathway (Eisenhut

et al. 2006) of phosphoglycolate metabolism in diatoms,

although genes required the glyoxylate cycle are present.

Enzymes in diatom peroxisomes include a glycolate

dehydrogenase that also oxidises L-lactate and enzymes of

the glyoxylate cycle (Winkler and Stabenau 1995).

Mitochondria have a glycolate dehydrogenase that also

oxidises D-lactate, and can convert the resulting glyoxy-

late to serine and CO2. Malate synthase in peroxisomes

can convert glyoxylate, with acetyl CoA which comes

ultimately from photosynthetically produced triose phos-

phate with CO2 release, into malate. Provided the rate of

phosphoglycolate synthesis is not too great, the products

of the PCOC up to serine hydroxymethyl transferase, and

of malate synthase, can all be used in biosynthesis of the

amino acids glycine and serine and those of the aspartate

and glutamate families, and as tetrapyrols, replacing

PEPC activity. The least energetically costly means of

metabolising any excess malate is conversion to triose

phosphate with further CO2 release, although complete

oxidation of malate to CO2 allows much of the energy put

in as reductant and ATP to be recouped as ATP (Raven

et al. 2000). Further work is needed on the relationship

between glycolate metabolism and N assimilation in

diatoms (Allen et al. 2006).

Conclusions

Despite recent advances, much work is required before we

have a clear understanding of the pathways of inorganic C

assimilation and related reactions in diatoms, and of how

Photosynth Res (2007) 93:79–88 85

123

very substantial inorganic C leakage is avoided (Granum

et al. 2005; Raven et al. 2007).

Acknowledgements Unpublished work reported here was per-

formed under grant NER/A/S/2001/01130 from the Natural Envi-

ronment Research Council, UK. We are grateful to Professor D G

Mann for pointing out the work of Schmid (2001), and to the com-

ments of two anonymous reviewers which have significantly im-

proved the paper.

References

Allen AE, Vardi A, Bowler C (2006) An ecological and evolutionary

context for integrating nitrogen metabolism and related signal-

ling pathways in marine diatoms. Curr Opinion Plant Biol

9:264–273

Apt KE, Zaslavkaia L, Lippmeier JC et al (2002) In vivo character-

ization of diatom multipartite plastid targeting signals. J Cell Sci

115:4061–4069

Armbrust EV, Berges JA, Bowler C et al (2004) The genome of the

diatom Thalassiosira pseudonana: ecology, evolution and

metabolism. Science 306:79–86

Badger MR, Andrews TJ, Whitney SM et al (1998) The diversity and

co-evolution of Rubisco, plastids, pyrenoids and chloroplast-

based CO2 concentrating mechanisms in algae. Can J Bot

76:1052–1071

Badger MR, Price GD, Long BM et al (2006) The environmental

plasticity and ecological genomics of the cyanobacterial CO2

concentrating mechanism. J Exp Bot 57:249–265

Baldauf SL (2003) The deep roots of eukaryotes. Science 300:1703–

1706

Banse K (1982) Cell volumes, maximal growth rates of unicellular

algae and ciliates, and the role of ciliates in the marine pelagial.

Limnol Oceanogr 27:1059–1071

Barker HA (1935) Photosynthesis in diatoms. Arch Mikrobiol 6:141–

156

Beardall J, Mukerji D, Glover HE et al (1976) The path of carbon in

photosynthesis by marine phytoplankton. J Phycol 12:409–417

Birmingham BC, Colman B (1979) Measurement of carbon dioxide

compensation points in freshwater algae. Plant Physiol 64:892–

895

Birmingham BC, Coleman JR, Colman B (1982) Measurement of

photorespiration in algae. Plant Physiol 69:259–262

Boldt R, Edner C, Kolukisaoglu U et al (2005) D-GLYCERATE 3-

KINASE, the last unknown enzyme in the photorespiratory cycle

in Arabidopsis, belongs to a novel kinase family. Plant Cell

17:2413–2420

Boyd C, Gradmann D (1999) Electrophysiology of the marine diatom

Coscinodiscus wailesii. I. Endogenous changes in the membrane

voltage and resistance. J Exp Bot 50:445–452

Burkhardt S, Amoroso G, Riebesell U et al (2001) CO2 and HCO3–

uptake in marine diatoms acclimated to different CO2 concen-

trations. Limnol Oceanogr 46:1378–1391

Burns BD, Beardall J (1987) Utilization of inorganic carbon by

marine microalgae. J Exp Mar Biol Ecol 107:75–86

Cassar N, Laws EA, Popp BN et al (2002) Sources of inorganic

carbon for photo-synthesis in a strain of Phaeodactylumtricornutum. Limnol Oceanogr 47:1192–1197

Colman B, Hosain ML (1980) The release of glycolate by a fresh-

water diatom. J Phycol 16:478–479

Colman B, Rotatore C (1995) Photosynthetic inorganic carbon uptake

and accumulation in two marine diatoms. Plant Cell Environ

18:919–924

Colman B, Huerta IE, Bhatti S et al (2002) The diversity of inorganic

carbon acquisition mechanisms in eukaryotic microalgae. Funct

Plant Biol 29:261–270

Coombs J, Volcani BE (1968) Studies on the biochemistry and fine

structure of silica shell formation in diatoms. Silicon-induced

metabolic transients in Navicula pelliculosa (Breb.) Hilse. Planta

80:264–279

Cox EH, McLendon GL, Morel FMM et al (2000) The active site

structure of Thalassiosira weissflogii carbonic anhydrase 1.

Biochemistry 39:12128–12130

Derelle E, Ferraz C, Rombauts S et al (2006) Genome analysis of the

smallest free-living eukaryote Ostreococcus tauri unveils many

unique features. Proc Natnl Acad Sci USA 103:11647–11652

Drum RW (1963) The cytoplasmic fine structure of the diatom,

Nitzschia palea. J Cell Biol 18:429–440

Drum RW, Pancratz HS (1964) Pyrenoids, raphes, and other fine

structure of diatoms. Am J Bot 51:405–418

Edwards GE, Franceschi VR, Vosnesenskaya EV (2004) Single-cell

C4 photosynthesis versus dual-cell (Kranz) paradigm. Annu Rev

Plant Biol 55:173–196

Eisenhut M, Kahlon S, Hasse D et al (2006) The plant-like C2

glycolate cycle and bacterial-like glycerate pathway cooperate in

phosphoglycolate metabolism in cyanobacteria. Plant Physiol

142: 333–342

Enami I, Suzuki T, Tada O et al (2005) Distribution of the extrinsic

proteins as a potential marker for the evolution of photosynthetic

oxygen-evolving photosystem II. FEBS J 272:5020–5030

Falkowski PG, Katz ME, Knoll AH et al (2004) The evolution of

modern eukaryotic phytoplankton. Science 305:354–360

Field CB, Behrenfeld MJ, Randerson JT et al (1998) Primary

production in the ecosphere: integrating terrestrial and oceanic

components. Science 281:337–360

Fielding AS, Turpin DH, Guy RD et al (1998) Influence of carbon

concentrating mechanism on carbon stable isotope discrimina-

tion by the marine diatom Thalassiosira pseudonana. Can J Bot

76:1098–1103

Geider RJ, Osborne BA, Raven JA (1985) Light dependence of

growth and photosynthesis in Phaeodactylum tricornutum (Bac-

illariophyceae). J Phycol 21:609–619

Geider RJ, Osborne BA, Raven JA (1986) Growth, photosynthesis

and maintenance metabolic cost in the diatom Phaeodactylumtricornutum at very low light levels. J Phycol 22:39–48

Giordano M, Beardall J, Raven JA (2005) CO2 concentrating

mechanisms in algae: mechanisms, environmental modulation,

and evolution. Annu Rev Plant Biol 56:99–131

Granum E, Myklestad SM (1999) Effects of NH4+ assimilation on dark

carbon fixation and b-1,3-glucan metabolism in the marine

diatom Skeletonema costatum (Bacillariophyceae). J Phycol

35:1191–1199

Granum E, Raven JA, Leegood RC (2005) How do marine diatoms fix

10 billion tonnes of carbon per year? Can J Bot 83:898–908

Hanson DT, Franklin LA, Samuelsson G et al (2003) The Chlamydo-monas reinhardtii cia3 mutant lacking a thylakoid lumen-

localized carbonic anhydrase is limited by CO2 supply to

Rubisco and not photosystem II in vivo. Plant Physiol 132:2267–

2275

Harada H, Nakajima K, Sakaue K et al (2006) CO2 sensing at ocean

surface mediated by cAMP in a marine diatom. Plant Physiol

142:1318–1328

Hillier W, McConnell I, Badger MR et al (2006) Quantitative

assessment of intrinsic carbonic anhydrase activity and capacity

for bicarbonate oxidation in photosystem II. Biochemistry

45:2094–2102

Holdsworth ES, Colbeck J (1976) The pattern of carbon fixation in the

unicellular alga Phaeodactylum tricornutum. Mar Biol 38:189–

199

86 Photosynth Res (2007) 93:79–88

123

Holtum JAM, Smith JAC, Neuhaus HE (2005) Intracellular transport

and pathways of carbon flow in plants with crassulacean acid

metabolism. Funct Plant Biol 32:429–449

Johnston AM, Raven JA (1996) Inorganic carbon accumulation by the

marine diatom Phaeodactylum tricornutum. Eur J Phycol

31:988–990

Johnston AM, Raven JA, Beardall J et al (2001) Photosynthesis in a

marine diatom. Nature 412:40–41

Kaczmarska I, Beaton M, Benoit AC et al (2006) Molecular

phylogeny of selected members of the order Thalassiosirales

(Bacillariophyta) and evolution of the fultoportula. J Phycol

42:121–138

Keeley JE, Rundel PW (2003) Evolution of CAM and C–4 carbon

concentrating mechanisms. Int J Plant Physiol 164:S55–S77

Kellog EA (1999) Phylogenetic aspects of the evolution of C4

photosynthesis. In: Sage RF, Monson RL (eds) C4 Plant biology.

Academic Press, San Diego, pp 411–444

Kerfeld CA, Saway MR, Tanaka S et al (2005) Protein structures

forming the shell of primitive bacterial organelles. Science

309:151–163

Kremer BP, Berks R (1978) Photosynthesis and carbon metabolism in

marine and freshwater diatoms. Z Pfl Physiol 87:149–165

Lane TW, Saito MA, Geroget GN et al (2005) A cadmium enzyme

from a marine diatom. Nature 435:42

Lee RE, Kugrens P (1998) Hypothesis: the ecological advantage of

chloroplast ER–the ability to outcompete at low dissolved CO2

concentrations. Protist 149:341–345

Lee RE, Kugrens P (2000) Ancient atmospheric CO2 and timing of

the evolution of secondary endosymbioses. Phycologia 39:167–

172

Lu YK, Theg SM, Stemler AJ (2005) Carbonic anhydrase activity of

the photosystem II OEC33 protein from pea. Plant Cell Physiol

46:1944–1953

Martin CL, Tortell PD (2006) Bicarbonate transport and extracellular

carbonic anhydrase capacity in Bering Sea phytoplankton

assemblages: results from isotope disequilibrium experiments.

Limnol Oceanogr 51:2111–2121

Matsuda Y, Hara T, Colman B (2001) Regulation of the induction of

bicarbonate uptake by dissolved CO2 in the marine diatom,

Phaeodactylum tricornutum. Plant Cell Environm 24:611–620

Matsuda Y, Satoh K, Harada H et al (2002) Regulation of the

expressions of HCO3– uptake and intracellular carbonic anhydr-

ase activity in response to CO2 concentration in the marine

diatom Phaeodactylum sp. Funct Plant Biol 29:279–287

Mayama S, Mayama N, Shihara-Ishikawa I (2004) Characterization

of linear-oblong pyrenoids with cp-DNA along their sides in

Nitzschia sigmoidea (Bacillariophyceae). Phycol Res 52:129–

139

Mitchell C, Beardall J (1976) Inorganic C uptake by an Antarctic sea-

ice diatom Nitzschia frigida. Polar Biology 16:95–99

Mitra M, Mason CB, Xiao Y et al (2005) The carbonic anhydrase

gene families of Chlamydomonas reinhardtii. Can J Bot 83:301–

308

Monson RK, Moore BD, Ku MSB et al (1986) Co-function of C3- and

C4-photosynthetic pathways in C3, C4 and C3 – C4 intermediate

Flaveria species. Planta 168:493–502

Montsant A, Jabbari K, Maheswari U et al (2005) Comparative

genomics of the pennate diatom Phaeodactylum tricornutum.

Plant Physiol 137:500–513

Morel FMM, Cox EH, Kraepiel AML et al (2002) Acquisition of

inorganic carbon by the marine diatom Thalassiosira weissflogii.Funct Plant Biol 28:301–308

Mortain-Bertrand A, Descolas-Gros C, Jupin H (1987) Short-term 14C

incorporation in Skeletonema costatum (Greville) Cleve (Bacil-

lariophyceae) as a function of light regime. Phycologia 26:262–

269

Mortain-Bertrand A, Descolas-Gros C, Jupin H (1998) Pathway of

dark inorganic carbon fixation in two species of diatoms—influ-

ence of light regime and regulatory factors on diel variations. J

Plankton Res 10:199–217

Myers J (1980) On the algae. In: Falkowski PG (ed) Primary

productivity in the sea. Plenum Press, New York, pp 1–15

Nimer NA, Brownlee C, Merrett MJ (1999) Extracellular carbonic

anhydrase facilitates carbon dioxide availability for photosyn-

thesis in the marine dinoflagellate Prorocentrum micans. Plant

Physiol 120:105–111

Park H, Song B, Morel FMM (2007) Diversity of cadmium-

containing carbonic anhydrase in marine diatoms and natural

waters. Env Microbiol 9:403–413

Parker MS, Armbrust EV, Piovia-Scott J et al (2004) Induction of

photorespiration by light in the centric diatom Thalassiosiraweissflogii (Bacillariophyceae): molecular characterization and

physiological consequences. J Phycol 40:557–567

Parker MS, Armbrust EV (2005) Synergistic effects of light,

temperature, and nitrogen source on transcription of genes for

carbon and nitrogen metabolism in the centric diatom Thalass-iosira pseudonana (Bacillariophyceae). J Phycol 41:1142–1153

Peers G, Price NM (2006) Copper-containing plastocyanin used for

electron transport by an oceanic diatom. Nature 441:341–344

Peltier G, Cournac L (2002) Chlororespiration. Annu Rev Plant Biol

53:523–555

Pronina NA, Semenenko VE (1992) Role of the pyrenoid in

concentration, generation, and fixation of CO2 in the chloroplasts

of microalgae. Soviet Plant Physiol 39:470–476

Raven JA (1984) Energetics and transport in aquatic plants. AR Liss,

New York

Raven JA (1987) The role of vacuoles. New Phytol 106:357–422

Raven JA (1995) Scaling the seas. Plant Cell Environ 18:1090–1100

Raven JA (1997) CO2 concentrating mechanisms: a direct role for

thylakoid lumen acidification? Plant Cell Environm 20:147–154

Raven JA, Spicer RA (1996) The evolution of crassulacean acid

metabolism. In: Winter K, Smith JAC (eds) Crassulacean acid

metabolism. Biochemistry, ecophysiology and evolution. Spring-

er, Berlin, pp 360–385

Raven JA, Waite AM (2004) The evolution of silicification in

diatoms: inescapable sinking and sinking as escape? New Phytol

162:45–61

Raven JA, Kubler JE, Beardall J (2000) Put out the light, and then put

out the light. J Mar Biol Assoc UK 80:1–25

Raven JA, Roberts K, Granum E et al (2007) The ecology and

evolution of single-cell C4-like photosynthesis in diatoms:

Relevance to C4 rice. In: Sheehy J, Mitchell P (eds) Charting

new pathways to C4 rice. World Scientific Publishers, Singapore

(in press)

Reinfelder JR, Kraepiel AML, Morel FMM (2000) Unicellular C4

photosynthesis in a marine diatom. Nature 407:969–999

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

Roberts SB, Lane TW, Morel FMM (1997) Carbonic anhydrase in the

marine diatom Thalassiosira weissflogii. J Phycol 33:845–850

Rost B, Riebesell U, Burkhardt S et al (2003) Carbon acquisition by

bloom-forming marine phytoplankton. Limnol Oceanogr 48:55–

67

Rotatore C, Colman B (1992) Active uptake of CO2 by the diatom

Navicula pelliculosa. J Exp Bot 43:571–576

Rotatore C, Colman B, Kuzuma M (1995) The active uptake of

carbon dioxide by the marine diatoms Phaeodactylum tricornu-tum and Cyclotella sp. Plant Cell Environ 18:913–918

Sage RF, Li M, Monson RK (1999) The taxonomic distribution of C4

photosynthesis. In: Sage RF, Monson RK (eds) C4 plant biology.

Academic Press, San Diego, pp 551–584

Photosynth Res (2007) 93:79–88 87

123

Satoh D, Hiraoka R, Colman B et al (2001) Physiological and

molecular biological characteristics on intracellular carbonic

anhydrase from the marine diatom Phaeodactylum tricornutum.

Plant Physiol 126:1459–1470

Schmid AM (2001) Value of pyrenoids in the systematics of diatoms:

their morphology and ultrastructure. In: Economou-Amilli A

(ed) Proceedings of the 16th international diatom symposium.

University of Athens Press, Athens, pp 1–32

Shiraiwa Y, Danbara A, Yoke K (2004) Characterization of highly

oxygen-sensitive photosynthesis in coccolithophorids. Jap J

Phycol 52 (suppl):87–94

Sicko-Goad L, Stoermer EGF, Ladewski BG (1977) A morphometric

method for correcting phytoplankton cell volume. Protoplasma

93:1–6

Sicko-Goad L, Stoermer EF (1979) A morphometric study of lead and

copper effects on Diatoma tenue var elongatum (Bacilla-

riophyta). J Phycol 15:316–321

Sicko-Goad L, Simmons MS, Lazinsky D et al (1988) Effect of light

cycle on diatom fatty acid composition and quantitative

morphology. J Phycol 24:1–7

Stemler AJ, Govindjee (1973) Bicarbonate ion as a critical factor in

photosynthetic O2 evolution. Plant Physiol 52:119–123

Strzepek RF, Harrison PJ (2004) Photosynthetic architecture differs in

coastal and oceanic diatoms. Nature 431:689–692

Szabo E, Colman B (2007) Isolation and characterization of carbonic

anhydrases from the marine diatom Phaeodactylum tricornutum.

Physiol Plant 129:484–492

Tanaka Y, Nakatsuma D, Harada H et al (2005) Localization of

soluble b-carbonic anhydrase in the marine diatom Phaeodacty-lum tricornutum. Sorting to the chloroplast and cluster formation

on the girdle lamella. Plant Physiol 138:207–217

Tcherkez GGB, Farquhar GD, Andrews TJ (2006) Despite slow

catalysis and confused substrate specificity, all ribulose bisphos-

phate carboxylases may be nearly perfectly optimised. Proc Natl

Acad Sci USA 103:7246–7251

Tchernov D, Helman Y, Keren N et al (2001) Passive entry of CO2

and its energy-driven intracellular conversion to HCO3– in

cyanobacteria are driven by a photosystem I-generated DlH+ . J

Biol Chem 276:23450–23455

Tsuji Y, Iwamoto K, Suzuki I et al (2006) A mechanism of

photosynthetic carbon fixation in the haptophyte alga, Emilianiahuxleyi. Plant Cell Physiol 47(suppl):s53

van Hunnik E, Sultemeyer D (2002) A possible role for carbonic

anhydrase in the lumen of chloroplasts in green algae. Funct

Plant Biol 29:243–249

Villareal TA, Joseph L, Brezinski MA et al (1999) Biological and

chemical characteristics of the giant diatom Ethmodiscus (Bac-

illariophyceae) in the Central North Pacific Gyre. J Phycol

35:896–902

Walker NA, Smith FA, Cathers IR (1980) Bicarbonate assimilation by

freshwater charophytes and higher plants. I Membrane transport

of bicarbonate ions is not proven. J Membr Biol 57:51–58

Whitney SM, Baldett O, Hudson GS et al (2001) Form I Rubiscos

from non-green algae are expressed abundantly but are not

assembled in tobacco chloroplasts. Plant J 26:535–547

Winkler U, Stabenau H (1995) Isolation and characterization of

peroxisomes from diatoms. Planta 195:403–407

Wittpoth C, Kroth PG, Weyrauch KV et al (1998) Functional

characterization of isolated plastids from two diatoms. Planta

206:79–95

88 Photosynth Res (2007) 93:79–88

123