Sie sind auf Seite 1von 14

Algal Research 1 (2012) 120133

Contents lists available at SciVerse ScienceDirect

Algal Research
journal homepage: www.elsevier.com/locate/algal

Review article

Bioremediation and other potential applications of coccolithophorid algae: A review


N.R. Moheimani , J.P. Webb, M.A. Borowitzka
Algae R&D Center, School of Biological Sciences and Biotechnology, Murdoch University, Murdoch, WA, 6150 Australia

a r t i c l e

i n f o

Article history:
Received 26 March 2012
Received in revised form 21 June 2012
Accepted 27 June 2012
Available online 24 July 2012
Keywords:
Calcication
Carbon concentrating mechanism
Pleurochrysis
Emiliania
Biofuel
Carbon models

a b s t r a c t
Coccolithophorid algae (Haptophycea) are mainly marine unicellular phytoplankton. The coccolithophorids
are of global interest as they can x carbon by photosynthesis as well as in calcium carbonate (coccoliths).
They are the largest carbon sinks and one of the largest primary producers on the planet. They can also produce high amounts of lipids which have a high potential application as a renewable fuel and alternative food
source. This paper reviews current knowledge on coccolithophorid algae photosynthesis and calcication and
their potential industrial applications.
2012 Elsevier B.V. All rights reserved.

Contents
1.
2.
3.
4.
5.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . .
Coccolithophorid algae . . . . . . . . . . . . . . . . . . . .
Ecology . . . . . . . . . . . . . . . . . . . . . . . . . . .
Calcication and photosynthesis . . . . . . . . . . . . . . .
Carbon concentrating mechanism (CCM) . . . . . . . . . . .
5.1.
Models of photosynthesiscalcication interactions . . .
6.
Ca2 + transport . . . . . . . . . . . . . . . . . . . . . . . .
7.
Effect of light on photosynthesis and calcication . . . . . . .
8.
Potential production of coccolithophorids . . . . . . . . . . .
9.
Other commercial applications of coccolithophids . . . . . . .
9.1.
Immobilisation of CO2 through biomass burial . . . . . .
9.2.
Biofuel . . . . . . . . . . . . . . . . . . . . . . . .
9.3.
Biomass coring . . . . . . . . . . . . . . . . . . . .
9.4.
The application of cell lipids and coccolithophorid algae as
9.5.
Applications of calcium carbonate . . . . . . . . . . .
10.
Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . .
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
pharmaceutical or
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .
. . . . . . . . .

. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
nutraceuticals
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

120
121
123
123
123
124
126
127
127
128
128
128
128
129
129
129
129
129

1. Introduction

Abbreviations: CCM, carbon concentrating mechanism; CDR, carbon dioxide removal;


CV, coccolith vesicle; Rubisco, ribulose-1,5-biphosphate carboxylase/oxygenase; CAext,
external carbonic anhydrase; AE, anion exchange.
Corresponding author. Tel.: +61 8 9360 2682.
E-mail address: n.moheimani@murdoch.edu.au (N.R. Moheimani).
2211-9264/$ see front matter 2012 Elsevier B.V. All rights reserved.
doi:10.1016/j.algal.2012.06.002

Before the Industrial Era atmospheric carbon dioxide concentration


was 280 ppm for several thousand years and it has risen continuously
since then, reaching 379 ppm in 2005 [1]. Potential global warming
induced by the accumulation of green house gases such as CO2 has become an important environmental issue [2]. CO2 is responsible for well

N.R. Moheimani et al. / Algal Research 1 (2012) 120133

over half the total warming potential of all greenhouse gases [3] and
annually about 20 billion tonnes of fossil CO2 are emitted from the
burning of fossil fuels, and another 2 to 8 billion tonnes are discharged
by human-mediated oxidation of the biosphere [4]. Increased levels of
CO2 will also have a signicant impact on oceanic pH. As atmospheric
carbon concentration increases, from present day to the forecasted
1200 ppm [5,6] the oceans will slowly become acidic which will result
in dissolving CaCO3. This can cause a negative feedback loop, resulting
in even more CO2 into the system, thus dropping the pH even further
[7]. This increase in atmospheric CO2 is one of the reasons for the
importance of investigations of methods for minimising and removing
anthropogenic CO2 emissions [8]. Some of the arguments for why
carbon dioxide removal (CDR) methods must be considered and developed are: a) the development of CDR methods has been proposed in
the United Nation Framework convention on climate change [8],
b) CDR could be a cost effective response in regards to high CO2 concentration and can be applied in the short and intermediate term [9],
c) application of CDR could be necessary in the longer term if other
mitigation methods fail and it is a possible low cost mitigation option
[8], and d) in case of sustainable fossil fuel use, CDR is the only green
house gas mitigation alternative [10].
Several methods have been proposed for capturing and removing
carbon dioxide: a) chemical or physical absorption processes which
are based on reactions between CO2 and one or more basic absorbents
such as aqueous solutions of sodium or potassium carbonate [11],
b) absorption processes which are based on signicant intermolecular
forces between gases and the surfaces of certain solid materials [11],
c) membrane processes which included either gas separation membranes or gas absorption membranes [11], d) absorbing CO2 from
multi-component gas streams and sequestering it in the deep ocean,
by injecting concentrated and liqueed CO2 into the deep sea or burying liqueed CO2 underground [1214], and e) bioxation of carbon
dioxide by photosynthetic organisms [10,15].
Photosynthetic organisms such as microalgae utilise solar energy to
x CO2 into organic carbon. Over the last few decades several studies
have emphasised the need to determine the potential of microalgal
cultivation systems to decrease CO2 emissions and to reduce or limit
the growing use of fossil fuels [2,12,16]. A signicant amount of carbon
dioxide is xed annually by plants [3], but based on the slow growth
rate of higher plants, the fresh water requirement, and the high cost
of land for growing these plants, this appears not to be a feasible option
for removing carbon dioxide. The cultivation of photosynthetic microorganisms such as algae and cyanobacteria has been proposed as
an alternative for CO2 bioremediation [2]. Algae are attractive organisms for CO2 bioremediation since they have a very high areal productivity when compared to other photosynthetic organisms such as
trees [10,17,18]. Microalgae cultures also have several characteristics
that argue for potentially higher productivities than higher plants
[10,19,20]: a) the possibility of culturing microalgae in continuous
large-scale systems, b) the ability to provide optimal nutrients levels
at all times, c) the absence of non-photosynthetic supporting structures, d) the ability to adjust harvest rates to keep the culture concentration at optimal levels at all times, e) the ability to control cell
composition (i.e. lipids) while increasing specic growth rate resulting
in higher productivities, f) the ability to grow at hyper saline condition
reducing the dependency to fresh water, and g) the ability to grow on
non agricultural land.
Due to favourable climatic and economic conditions, some countries offer numerous relatively large and low cost opportunities for
indirect biological CO2 mitigation, through forestry, agriculture, and
biofuel projects [10]. However, several technical problems remain to
be resolved, including long-term feasibility and transactional costs
[21]. The problem of disposal of the large volume of captured CO2
also is still unsolved [22]. For a CO2-removal process to be both technically and economically feasible, the development of a well-engineered
system that converts CO2 to useful products, would be advantageous

121

[23]. When it comes to the use of microalgae for CO2 bioremediation,


coccolithophorid algae are of interest, mainly due to their ability to
form CaCO3 scales together with photosynthetic carbon xation. The
carbon xed by photosynthesis is a part of the carbon cycle, while
the CaCO3 can be considered to be discharged (precipitated) out of
the carbon cycle. Bioremediation of CO2 and other potential applications of coccolithophorids are the main focus of this review.
2. Coccolithophorid algae
The coccolithophorid algae are unicellular agellates containing
chlorophyll a and c and are classied as members of the Haptophyceae.
As cited in Green and Jordan [24], Ehrenberg made the rst recorded observation of coccoliths in 1836, while examining chalk from the island of
Rugen in the Baltic Sea. During the last 70 years, especially since the introduction of the electron microscope, there has been a massive increase
in knowledge regarding the diversity of microalgae and especially the
Haptophyceae [25]. Towards the end of the nineteenth century the
importance of coccolithophorids and other calcareous nanoplanktonic
organisms as a primary link in the marine food chain was recognised,
and this stimulated the study of the biology of these organisms [25].
In more recent years, the study of coccolith samples is an important
tool for micropalaeontological studies for industrial laboratories due
to the associations of coccolithophorids with fossil oil production
[26].
The Haptophyceae are mainly marine, but a few freshwater and
terrestrial species are also known [24]. The most abundant representatives of the Haptophyceae are the coccolithophorid algae. They are
unicellular, can be either motile or nonmotile, with the ability to produce external calcite plates (coccoliths) which cover their surface [25].
In the calcied coccolithophorids, the cell surface is covered by one
(e.g. Pleurochrysis carterae) or several layers (e.g. Emiliania huxleyi) of
coccoliths. Coccoliths are mostly oval-shaped structures of about 1 to
20 m in diameter consisting of some organic material and crystalline
calcite CaCO3.
Coccolithophorids are normally classied as either heterococcolithophorids or holococcolithophorids [27,28]. Heterococcoliths
are complex multi-component structures, while holoccoliths are simple
single crystalline structures [29]. Interestingly, culture studies showed
that most holococcolithophorid algae are in the haploid phase [29,30]
and can be a part of heterococcolithophorid life cycle, however this
hypothesis has been tested only in some species such as Coccolithus
pelagicus, Coccolithus haylinus, Calcidiscus leptoporous, Coronosphaera
mediterranea and E. huxleyi [31,32]. Most coccolithophorids are photosynthetic although Balaniger balticus is non-photosynthetic (obligate
heterotrph) [24].
Many coccolithophorids such as P. carterae are motile or produce
motile cells as a part of their life cycle [33]. Coccolithophorids with
agella usually have two subequal to unequal agellae with no
hair-like appendages [34]. The agellar apparatus has been demonstrated to be of phylogenetic importance in that it is considered to
be an evolutionary conservative feature [35].There is a wide diversity
in the life cycle of coccolithophorid algae. Many of the genera are yet
to have their life cycles reported, and as yet the only detailed information is on a few of the genera such as Emiliania and Pleurochrysis
[36,37]. Life cycle of coccolithophorid algae such as E. huxleyi is mostly
asexual [37]; however, many (such as Pleurochrysis) have an alteration of generation [30,38,39]. At present three morphotypes of
Pleurochrysis have been identied, the diploid heterococcolith stage,
the haploid holococcolith stage (Apistonema), which has a covering
of organic scales [30,33,39,40] and the diploid pseudolamentous
benthic stage [39]. The main four life cycle stages of Pleurochrysis
are: a) diploid motile coccolith bearing stage [41], b) colonial non
motile coccolith bearing stage [38,39], c) haploid benthic apistonema
stage (haploid) [33,42] and d) haploid benthic lamentous stage
[39,43]. The motile coccolith bearing stage generally reproduces

122

N.R. Moheimani et al. / Algal Research 1 (2012) 120133

by a sexual cell division [39], however, diploid coccolith bearing


Pleurochrysis may also produce a colonial stage consisting of 24
cells with an outer covering of coccoliths [38]. There are also two
motile, non-coccolith stages have been observed, swarmers produced
by the haploid stage and gametes. While the heteromorphic life cycle
has only been described in detail in Pleurochrysis psuedoroscoffensis
[39], it can be assumed that it holds true for P. carterae, as the differences in the haploid and diploid scale covering have also been seen
in P. placolithoides and P. lacuna [30,44]. It is to be noted that in
E. huxleyi, and most other coccolithophorids, alterations of generations have not been observed. However, variation in morphology of
haploid and diploid of E. huxleyi has been reported [37].
One of the interesting characteristics of the Haptophyceae is the
presence of a haptonema, originally reported by Manton [45]. The
haptonema is a highly variable organelle [46]. The main function of
the haptonema is still unknown, however, at this stage it seems to
be an organelle with various functions such as: a) attachment to
other cells, b) as a sensory and tactile organelle and/or c) for capturing other planktonic organisms and playing a role in phagocytotic activity [35]. The existence of the haptonema in coccolithophorids and
its possible function in feeding on marine bacteria also indicate a
mixotrophic life style [46]. For example, some coccolithophorid
algae such as P. carterae have been shown to be heterotrophic at
some stages during the life cycle and utilise the relatively high concentrations of dissolved organic molecules present in the photic
zone [4751].
While the main reason for biomineralisation in coccolithophorid
algae remains unsolved, several hypotheses have been proposed by
investigators [36,5254] for coccolith formation: a) focusing light
onto the chloroplast, b) light reection to decrease photoinhibition,
c) augmentation of the internal supply of photosynthetic CO2, d) increasing the cell volume, e) protecting the cell, and f) increasing the
sinking rate. There are some very good biochemical and evolutionally
arguments for coccolithophorid calcication, however, none have yet
to explain why this genus produced such ornate and complex calcite

structures, and as such, , the main function of coccoliths has yet to be


resolved.
Coccolith and the cell structure of two members of the
coccolithophorid algae (E. huxleyi and P. carterae) have been extensively
studied, mainly because they are easily grown in culture [25,55,56]. The
morphology of both of these species has been well described in previous
studies [5761]. In both of these species coccoliths are produced
intracellularly in a coccolith vesicle (CV) and they are then extruded
from cell and placed into the coccosphere [61] (Fig. 1). In E. huxleyi
(heterococcolithophorid) electron microscope studies indicate that
the CV is located beside the nucleus enclosing the growing coccolith
and is connected to the reticular body (Fig. 1). The coccoliths are
composed of a radial array of unit elements in which one unit is a single
crystal of calcite [62]. In a completed coccolith these crystals have
rounded faces and edges [29]. Prior to calcication the coccolith
vesicle/reticular body (CVRB) system appears to contain a high concentration of acidic polysaccharides relative to other cell constituents
[63]. It has been hypothesised that this polysaccharide plays a regulatory role in the calcication process [61,64]. Calcication in E. huxleyi is a
highly controlled process and initiated at the base of the future
connecting wall of the coccolith and progresses rst upward and then
perpendicularly to form the wall and upper element of the coccolith,
respectively. It then progresses outward and inward at the base to
form the lower element (Fig. 1) [62,63]. Completed coccoliths are then
extruded and incorporated into the coccosphere [65]. Each E. huxleyi
cell is surrounded by about 1520 coccoliths and this species is capable
of releasing coccoliths to the medium (Fig. 1). In this species it has
been found that, when external coccoliths are removed from cells by
acidication, the formation of a new coccosphere will be initiated [66].
In holococcolithophorid algae such as Poricalyptra aurisinae, some studies suggest that the trans-Golgi apparatus appears to be a coccolith storage and production compartment. However, Young et al. [29] state that
intercellular holococcolith calcication has never been observed. The
lack of information on coccolith structure and formation in other species
of coccolithophorids leaves this subject open to further investigation.

Fig. 1. Diagrams of cross sections through calcifying cells of Emiliania huxleyi (from van der Wal et al. [211]) and a through a cell of Pleurochrysis carterae (from van der Wal et al.
[63]). While the E. huxleyi cell is covered by a multilayer coccolith layer (only two of the extracellular coccoliths (EC) in this gure), the Pleurochrysis carterae cell is covered by a
single layer of EC followed by several layers of organic scales (sc), and extreme proximally by columnar material (CM) [212]. In E. huxleyi, coccoliths are composed of about 30 units
of radially arranged crystalline units each of which can be subdivided into a connecting wall (a) between lower element (b) and an upper element (c). Different intracellular
organelles of both species are also shown including chromatin (Ch); chloroplast (Chl) cover (Cov); Golgi complex (G); mitochondrion (M); nucleus (N); nuclear envelope (NE);
reticular body (RB); cell vacuole (V); crystalline matter (X) autophagic vacuole (AV), endoplasmic reticulum (ER), coccolith vesicle (CV), and coccolithosomes (cs). The ve
morphologically discernible stages of the coccolith production compartment of E. huxleyi are also shown in the right [62].

N.R. Moheimani et al. / Algal Research 1 (2012) 120133

3. Ecology
Coccolithophorids are widely distributed throughout the oceans,
and in some areas they may become the dominant members of the phytoplankton community [67]. Certain species, such as E. huxleyi can form
massive seasonal blooms which are visible from space, due to light
scattering by the coccoliths [6870]. Blooms of coccolithophorid algae
such as Pleurochrysis pseudoroscoffensis and P. carterae have also been
reported in inland saline lakes such as the Salton Sea, California [71].
The highest diversity of coccolithophorids today is in the subtropical
oceanic gyres [72] and coccolithophorid diversity is much lower in
polar waters than in the tropics [73]. The coccolithophorid algae,
together with foraminiferans, represent the bulk of modern-day
global CaCO3 production [57]. It is estimated that the coccolithophorid
algae represent up to half of all existing global CaCO3 production
[74,75]. Of the 195 coccolithophorid taxa, E. huxleyi is the most
abundant coccolithophorid living in today's oceans and is a very abundant phytoplankton [54]. The sedimentary records indicate the dominance of E. huxleyi over the rest of coccolithophorids during the last
73,000 years [76].
Coccolithophorid algae are not fully photoautotrophic as they have
a requirement for one or more exogenous sources of vitamins and in at
least one species phagotrophy has been observed [77,78]. Many studies
have shown that low phosphate and nitrate concentrations induce
coccolith production in some coccolithophorids such as E. huxleyi and
P. carterae [7982]. Studies in the open ocean have shown that the
highest cell densities of coccolithophorids under non bloom conditions
are observed at the lowest nitrogen and phosphorus conditions [83,84].
Moreover, Shiraiwa [85] found a signicant increase in coccolith
production under high bicarbonate, low phosphate and low nitrate
concentrations in E. huxleyi, whereas an increase in bicarbonate
concentration resulted in an increase in growth and suppression of
calcication to the point that the cells lost their coccolith layers
(Fig. 2). It has also been shown that oceanic coccolithophorids such
as E. huxleyi need less iron, zinc and manganese compared to coastal
species such as P. carterae [86]. These data indicate the possible
evolution of the mechanism(s) that allows oceanic coccolithophorids

123

to survive in low nutrient conditions. For example, in the case of iron,


it appears that oceanic coccolithophorids achieved this ability by
reducing the required amount of iron per cell [87]. It has also been proposed that the calcication response to low nutrients is due to the activation of the plasma membrane transporter(s) that exchange protons
for nutrient ions [88]. The high competitiveness of coccolithophorids
in the oceanic environment can also be attributed to a possible antibiotic secretion by these algae. The fact that microalgae may produce
antibiotics has been known for some time [89]. Some species of
coccolithophorid algae have been found to inhibit the growth of
other phytoplankton species by producing toxins [90] and some clones
of Pleurochrysis pseudoroscoffensis have been found to release substances (unknown) which are toxic to other phytoplankton, whereas
laboratory experiments indicated no toxic effect of these substances
on mice and brine shrimp [91].
4. Calcication and photosynthesis
Calcium carbonate precipitation is a common phenomenon in
seawater, fresh water and soil environments [92]. It is a rather straightforward chemical process governed by four key factors including
calcium concentration, the concentration of inorganic carbon (Ci) the
pH, and the availability of nucleation sites [52,9395]. Coccolithophorids
x inorganic carbon via photosynthesis and intracellular calcication. Calcication and photosynthesis have been shown to be closely
linked in many calcied algae [95]. Photosynthesis and calcication
in coccolithophorids with particular emphasis on E. huxleyi have been
widely reviewed [35,57,63,96]. The various aspects of photosynthesis
and calcication of coccolithophorids are presented in detail below.
5. Carbon concentrating mechanism (CCM)
Marine phytoplankton are potentially CO2 limited because of the
physico-chemical properties of seawater. Less than 1% of dissolved
inorganic carbon (DIC) is present as CO2 at normal seawater pH
(8.18.2) and more than 90% occurs in the form of HCO3 (Eq. (1))
[97]. CO2 diffusion from air to seawater is low, followed by a low

Fig. 2. A schematic diagram for regulation of Emiliania huxleyi cell growth and calcication by nutrient supply in oceanic conditions.
From Shiraiawa et al. [124].

124

N.R. Moheimani et al. / Algal Research 1 (2012) 120133

conversion rate of bicarbonate to carbon dioxide at normal seawater


pH (8.1) and this contributes to the low availability of CO2 to marine
photosynthetic organisms [98].

CO2 H2 O

pk1 6:3

HCO3

pk2 10:25

CO3

2H

The general bicarbonatecarbonate buffer system can also provide


CO2 for photosynthesis. The OH produced in the conversion of bicarbonate to CO2 (Eq. (2)) can lead to a gradual rise in the medium pH
[99102]. It is common to measure a pH as high as 11 in high density
algal production systems where no additional CO2 has been supplied
[103,104].

HCO3 CO2 OH

The limitation in the carbon dioxide supply can restrict carbon assimilation in marine microalgae since they rely on CO2 for photosynthesis.
Many phytoplankton species have been shown to have evolved a
carbon dioxide concentrating mechanism (CCM), which permits them
to use either CO2 or HCO3, or both, as external sources of inorganic carbon (Ci) [105107]. The CCM mechanisms makes it possible for cells to
enhance the delivery of CO2 to ribulose-1,5-biphosphate carboxylase/
oxygenase (Rubisco) and limit the oxygenase activity of this enzyme
[108]. Calcifying microalgae also use Ci to produce calcium carbonate
and early studies have shown that most calcifying algae can use HCO3
as the carbon source for calcication [95,109]. Based on the low afnity
of Rubisco for CO2, on the other hand, the possibility of photosynthetic
dependence of calcifying algae on calcication has been suggested and
investigated in a number of studies [53,80,110,111]. The use of bicarbonate as the substrate for calcication results in the net production
of CO2 during calcication (Eq. (1)) [75,112]. The fate of this CO2 in photosynthesis is still unknown but, if bicarbonate is the external carbon
source for both calcication and photosynthesis, then the H+ released
during calcication (Eq. (1)) can be neutralised by OH- produced during
photosynthetic CO2 uptake (Eq. (2)) [99,113]. The potential advantage
of calcication as an energy-efcient way of supplying Rubisco with
CO2 in E. huxleyi has also been suggested by [114]. Studies of CCMs
have concentrated on cyanobacteria and freshwater green algae and almost invariably involve the active transport of the species CO2, H+ and
HCO3 across one or more membranes in bringing about a higher
steady-state concentration of CO2 available to Rubisco than presents
in the bulk medium [108,115]. In these organisms, it has been shown
that the main components of the CCM are those which cause carbon accumulation in the cell or chloroplast [116]. The CCM functions either by
actively transporting CO2, HCO3 or both, and/or by the contribution of
an external carbonic anhydrase (CAext) which catalyses the conversion
of the two forms of Ci [117].
Buitenhuis and colleagues [74] showed that over 90% of carbon
used by E. huxleyi in photosynthesis came from bicarbonate. Sikes
and Wheeler [118] have also described the use of bicarbonate by
E. huxleyi in photosynthesis under high alkaline conditions. Although
the existence of a CAext in E. huxleyi has been reported in many
studies, the activity of this enzyme at normal seawater pH (8.1)
has been a mystery for a long time and has been the subject of
many investigations [118120]. Herfort et al. [121] demonstrated
the existence of a membrane anion exchange (AE) protein and a
CAext in E. huxleyi, both of which are involved in active bicarbonate
transport into the cell. They [121] also showed that in E. huxleyi, CAext
is only active at low Ci concentrations (b0.5 mM), resulting in no activity of this enzyme at the normal Ci (=2 mM) concentration and pH
(8.1) of seawater. On the other hand, they showed that the AE was active at all levels of Ci. Furthermore, CAext appears to have no obligate
role in Ci uptake by E. huxleyi and could therefore have evolved to enable competition with other phytoplankton at the low Ci concentration

uptake in the ocean (e.g., in conditions after an algal bloom). While, the
CAext is not typically active at normal Ci condition in E. huxleyi, Israel and
Gonzales [122] have demonstrated activity of CAext at both high and low
Ci concentrations in Pleurochrysis sp. No CAext activity, however, was
detected by Huertas and colleagues [117] in the heterococcolithophorid
Ochrosphaera neopolitana. These contradictory results suggest that
there is a high possibility of various CCMs between different species of
coccolithophorid algae.
5.1. Models of photosynthesiscalcication interactions
Over the last 60 years there have been many attempts to develop a
conceptual model for calcicationphotosynthesis in coccolithophorid
algae, with the main emphasis on E. huxleyi [53,57,96,123,124]. One
of the earliest models for the interaction between photosynthesis and
calcication of E. huxleyi was developed by Paasche [123] assuming
bicarbonate usage for both photosynthesis and calcication. In his
model photosynthesis and calcication were not linked closely, mainly
due to the results achieved from the clone of E. huxleyi studied
including: a) different light saturation kinetics of photosynthesis
and calcication, b) the same rate of photosynthesis between cells
grown in calcium-free and calcium-rich medium, and c) less effect of
photosystem II inhibitors on calcication when compared to photosynthesis. However, he did not indicate any close association between
photosynthesis and calcication. Later studies, mainly based on the
calcicationphotosynthesis (C:P) ratio of other clones of E. huxleyi,
indicated that while not all CO2 produced in calcication can be utilised
in photosynthesis, some of the released CO2 during calcication can
be used in photosynthesis [61,74,118]. Paasche [123] also assumed
the active transport of protons and hydroxyl ions produced during
calcication and photosynthesis from cells of E. huxleyi to the medium,
however no data has been found in coccolithophorids supporting this
hypothesis.
Moreover, Sikes and colleagues [53] subsequently showed that the
H + produced by HCO3- conversion to CaCO3 may be used to produce
an extra CO2 from the second HCO3 taken up. It should be noted
that this model did not indicate coccolith vesicle as the actual site
for HCO3 CO2 conversion whereas many studies since have
indicated the coccolith vesicle as the site of calcication [29,79,125].
They [53] also proposed a number of possible alternatives for Ci diffusion mechanisms for photosynthesiscalcication in coccolithophorids
(Fig. 3). Their results showed that, for E. huxleyi, the equilibrium
showing in Fig. 3B seems to be the most likely Ci uptake mechanism,
however they could not reject the possibility of the use of carbonate
for calcium carbonate production (Fig. 3C). In E. huxleyi scheme A, in
Fig. 3, is the least likely in that it invokes CO2 as the source of Ci for
coccoliths, since, other studies on E. huxleyi [111,120] have also
shown that this species denitely uses bicarbonate for calcication.
Later on Brownlee et al. [96] developed a more detailed method for
calcication and photosynthesis in E. huxleyi. In this model the partial
reactions of calcication and the hydration /dehydration of CO2 occur
in separate compartments such as the coccolith vesicle, chloroplast or
cytoplasm of the cell. This could explain the isotopic disequilibrium
results which could be mainly due to H+ limitation (decreasing the
pH) on calcication process [53]. The possibility of using extra CO2 produced by calcication has also been considered. However, this was the
rst time that the possibility of extra CO2 excretion to the medium was
reported.
Although, all models and results indicated the usage of bicarbonate
in E. huxleyi, McConnaughey [88] suggested CO2 as the source of Ci for
both calcication and photosynthesis mainly based on the amount of
H+ required for balancing the charge produced by Ca 2+ in the cell.
However, this model was not supported by the results of Buitenhuis
and colleagues [74] who demonstrated a high bicarbonate requirement
for photosynthesis and calcication in E. huxleyi. Nevertheless, the
McConnaughey [88] model might be a reasonable model for other

N.R. Moheimani et al. / Algal Research 1 (2012) 120133

125

Fig. 3. Alternative interpretations of various Ci sources supplied to the cell for photosynthesis and calcication (A) usage of only CO2 in both reactions, (B) usage of bicarbonate
in calcication and CO2 in photosynthesis, and (C) usage of carbonate in calcication and
CO2 in photosynthesis.
Redrawn from [53].

species of coccolithophorid algae such as P. carterae where no active


bicarbonate transport system such as AE1 protein has yet been
detected. Israel and Gonzales [122] found that CO2 was the sole source
of Ci in photosynthesis in low and high calcifying species of Pleurochrysis
sp. Additionally, their results also demonstrated the use of HCO3 in
calcication in high calcifying species, while the Ci source for calcication in low calcifying P. carterae has yet to be found [122].
The model presented in Fig. 4 is a summary of all investigations on
calcication and photosynthesis and is the most accepted model
available for E. huxleyi [74,96,114,126]. In this model, HCO3 is the
source for both calcication and photosynthesis, resulting in perfect
proton and hydroxyl ion balances maintaining a pH of approximately
7 in the cytoplasm. This model also allows CO2 to be exported from
the cell and results in a balance of the amount of up to 3 HCO3imported into the cell. Excretion of this CO2 from E. huxleyi cell can
also explain the reports regarding the increase in the atmospheric
CO2 by some blooms of E. huxleyi [93,127,128]. This CO2 release can
also explain the pH drop in actively growing E. huxleyi cultures
[100]. However, a very high pH increase (up to pH 11) found in the
outdoor culture of P. carterae [100,102]. An almost 1020 fold higher
reported C:P ratios for E. huxleyi (0.51.5) compared to P. carterae
(0.030.07) can result in a higher excretion of H + or CO2 and may
explain the difference between medium pH between P. carterae and
E. huxleyi [102]. Raven's [98] version of model suggests the possible
interaction between H+ and Ca2+ transportation into the coccolith
vesicle explaining the production of CO2 from HCO3- in the plastid
from H+ produced in the coccolith vesicle by possible carbonic
anhydrase activity (Fig. 5). This model also indicates the pH, Ci concentrations, and electrical potential in the seawater, coccolith vesicle and
plastid of E. huxleyi (Table 1). This model estimates that ATP-driven
Ca2+:H+ antiport in endomembranes are consistent with 1 Ca2+ and
1 HCO3 entering the coccolith vesicle and 1 H+ exiting [98]. Data
presented in Table 1 also indicates that a carbonic anhydrase seems
to be located in the chloroplast of the coccolithophorids and this
would seem to be the site of CO2 production. The pH has been found
to be lower in the coccolith vesicle than in seawater [114]. This lower
pH, together with the negative electrical potential of the coccolith

Fig. 4. A schematic diagram showing the putative transport routes of Ca2+ to the site of
calcication in the Golgi/CV. (A) Entry across the plasma membrane and diffusion
across the cytoplasm coupled with active pumping of Ca2+ into Golgi and/or CV.
(B) Endocytotic uptake of Ca2+ and eventual delivery of Ca2+ to the Golgi/CV via vesicle
transport. (C) Entry of Ca2+ into, and transcellular transport through, the cortical endoplasmic reticulum (CER), followed by ER-Golgi vesicle transport.
Redrawn from [125].

vesicle, indicates that carbonic anhydrase (i.e. internal CA) is unlikely


to exist at the site of calcication, whereas there is a high possibility
for CA existence at the site of photosynthesis in the coccolithophorids
(Fig. 5). As previously described, despite the high possibility of bicarbonate conversion to CO2 in the plastid by internal carbonic anhydrase,
free CO2 especially in seawater with a pH lower than pH 8.1, is the most
likely resource of Ci for photosynthesis. The possibility of direct CO2
uptake has been largely ignored in many of the most recent bicarbonate models for calcication and photosynthesis in coccolithophorids
[98,114]. While in E. huxleyi bicarbonate may be the sole source of
calcication and photosynthesis, CAext must be the only functioning
inorganic carbon transport mechanism for bicarbonate utilisation at
pH higher than pH 8.1. Under such conditions, CO2 has to be the
main resource for calcication and photosynthesis and the model
presented by McConnaughey in [88] seems to be the most likely
model for calcication and photosynthesis in these species. Furthermore, the McConnaughey model can also be supported by the observation of Crenshaw [129] who found that the HCO3- used in calcication in
P. carterae were produced from the CO2 derived from photosynthesis.
In contrast, Israel and Gonzalez [122], studying the effects of external
pH on calcication and photosynthesis in a high calcifying strain of
Pleurochrysis, showed that an adequate CO2 supply to photosynthesis
required bicarbonate utilisation and calcication. On the other hand,
it is not known that Crenshaw [129] used high or low calcifying strain
of P. carterae. These contradicting results specically indicate the need
for more research on the issue of inorganic carbon source for calcication in Pleurochrysis. On the other hand, subsequent studies such as
[57,121,130] indicated that at least in E. huxleyi it is external bicarbonate which is used in calcication. However, it is yet to be known what

126

N.R. Moheimani et al. / Algal Research 1 (2012) 120133

Fig. 5. The most accepted model for photosynthesiscoccolith formation presented initially by [126] reproduced by [114], upgraded by [98] and also in agreement with the results of
[74]. Empty circles are the places with energy requirement for ion transport.

species of C is used for calcication in other coccolithophorid. Moreover, the results of Buitenhuis et al. [74] for E. huxleyi also support the
model presented in Fig. 5. The main difference between all current
models can be explained by the calcication regulation of cells under
different conditions.
However, although current models encompass and predict most aspects of calcication and photosynthesis of coccolithophorids (mainly
in regard to E. huxleyi), some data from the literature cannot as
yet be fully integrated into any existing models. There are few
reports indicating calcication of coccolithophorids in the dark and
such a phenomenon has yet to be resolved [126]. Unfortunately,
the limited data available for species other than E. huxleyi does not
permit the development of a universal model for all coccolithophorid
algae. For example, as previously mentioned, the lack of active bicarbonate transporters in some species may require a new model for
photosynthesis and calcication.
In summary, calcication, photosynthesis, membrane transport
and metabolism appear to interact in a complex manner. We are
only beginning to understand and quantify the components of this
complex system. For example, Herfort et al. [121] have shown that
photosynthesis does not necessarily depend actively on calcication
as they did not detect any signicant differences in photosynthesis between a coccolith producing and a non-coccolith producing
E. huxleyi clone. Moreover, until now most studies have been carried
on E. huxleyi, while it is quite possible that the calcication and

Table 1
The pH and electrical potential of E. huxleyi at different sites [98].

Seawater
Plastid
Coccolith vesicle

pH

[HCO3]
mol m3

[CO2]
mmol m3

Electrical potential
(mV)

8.0
7.0
7.2

0.2
0.2
0.2

10
10
10

0
60
66

photosynthesis mechanisms are different between this species and


other coccolithophorids such as P. carterae. For example, the differences between CAext activity between E. huxleyi and P. carterae may
result in different balances between Ci sources for calcication and/
or photosynthesis. As previously mentioned the lack of information
and research into other species of coccolithophorids, especially in
the area of CCMs, leaves calcication/photosynthesis open to further
future investigations.
6. Ca 2+ transport
An understanding of the pathways and mechanisms of Ca 2+
transport is essential for identifying the mechanism and role of
calcication. Ca 2+ is a micronutrient in a number of microalgae
species and relatively little is known of its transport into the cell
[131]. An ATP-powered Ca 2+ pump is found in metazoa, protozoa,
plants and prokaryotes [131]. Ca 2+ ATPase also appears to be associated with calcication in various animals and plants [65,132134].
Calcication in coccolithophorid algae is dependent on Ca 2+ uptake
from the external medium to the coccolith vesicle (CV) [131]. While
the cytoplasm free calcium concentration of E. huxleyi is very low
[96], a high level of total Ca 2+ has been shown in the coccolith vesicle
of this species [63]. In E. huxleyi the cytoplasmic concentration of
Ca 2 + appears to be ve times lower than in the CV [30], therefore a
concentrating mechanism for Ca 2+ must exist in this species. All of
the calcium carbonate which ends up in the coccolith has passed
through cell and Golgi membranes. The energy required for active
Ca 2 + transport can be up to 20% of the total cost of xing an equivalent amount of carbon in the Calvin cycle [114] and light dependency
of Ca 2+ accumulation in P. carterae has been demonstrated [63]. Ca 2+
transfer is found to be via the formation of coccolithosomes, and the
number of coccolithosomes required to build one coccolith has been
calculated to be 4.5 [63]. Ca 2 + can enter the cell by diffusion
(Fig. 4A) but requires an ATP for active transport into the CV. Ca 2+
transport can also take place by formation of membrane packets at

N.R. Moheimani et al. / Algal Research 1 (2012) 120133

the cell membrane (Fig. 4B). These packets eventually combine with
the Golgi apparatus [30,114]. If, however, the distance required for
Ca 2+ penetration into the cortical endoplasmic reticulum (CER) is
small the ER body may have a role in Ca 2+ transport (Fig. 4C). In
E. huxleyi and P. carterae, Ca 2+ is found to be concentrated in the
Golgi apparatus [63,135], while there are no data for other species
of coccolithophorids.
7. Effect of light on photosynthesis and calcication
Coccolithophorid algae have several light harvesting pigments including chlorophyll a and c, -carotene, fucoxanthin, diatoxanthin
and diadinoxanthin, similar to diatoms and chrysophytes [136]. The
irradiance required for coccolithophorid growth and to saturate
photosynthesis also does not appear to be signicantly different from
that of other eukaryotic phytoplankton [137]. There are, however, species differences, with E. huxleyi and P. carterae requiring a vefold
higher irradiance compared with Gephyrocapsa oceanica [137]. It has
also been shown that high irradiances do not inhibit the growth of
E. huxleyi [137], however, this species has been found to be sensitive
to ultraviolet radiation [138,139]. Blooms of E. huxleyi are found to be
induced by high irradiance [57]. Emiliania huxleyi tolerance to high irradiance, even up to 1700 to 2500 mol photons m2 s 1, may be
explained by high light reection and the light scattering ability of
coccoliths [70,140,141].
Algal calcication is highly inuenced by light (see reviews [95]
and [126]). In one of the earliest studies Paasche [142] observed the
maximum calcication rate of E. huxleyi at the blue end of the
spectrum and he hypothesised this as the possible stimulator for
active bicarbonate transport. There is a wide body of information on
the dependence of calcication on irradiance and photosynthesis in
coccolithophorids, especially in E. huxleyi [53,57,80,110,111], however calcication in E. huxleyi has been found to be light saturated at
lower irradiances than photosynthesis [123]. At night, on the other
hand, decalcication or low calcication rates have been observed
in E. huxleyi [109,110,143]. A number of other studies also indicate
that calcication and coccolith formation in E. huxleyi is a light
dependent process [66,123,144]. Van der Wal and colleagues [145],
however, found the same rate of calcication, in P. carterae, during
both dark and light cycles. All of these studies show that calcication
appears to be light dependent in some species of coccolithophorids
such as E. huxleyi, but there may be some exceptions.
8. Potential production of coccolithophorids
Since a United Nations committee recommended that conventional
agriculture be supplemented with high-protein foods of unconventional origin, microalgae have become natural candidates for this [146]. On
the other hand, microalgae are attractive potential sources for sustainable biofuel production [20,147]. The primary source of all food and organic raw materials is solar energy [148150]. Exponential increases in
the world's population and its demands for nding possible resources
of food and energy will depend on how efciently we can learn to
use solar energy. Conventional agricultural systems are very inefcient
in this respect as a) most plants can only utilise less than 0.5% of the
sun light that falls on them, b) most farms cover only a small land
area, c) only a small proportion of each crop plant is edible, and d)
maximal production is highly limited by the availability of CO2 [148].
Microalgae promise important advantages to improve the solar
efciency utilisation as: a) they can be grown in continuous culture
providing maximal annual productivity, b) microalgal cells contain
relatively low structural material with the possibility of using
the whole biomass for nutrition or other economic uses, and c) addition of CO2 to a microalgal culture systems is relatively simple
compared to eld crops [20,102,147,151154]. Despite all of these advantages of microalgae over conventional agriculture, the feasibility of

127

microalgae as food or energy sources has yet to be proven and it


is mainly limited by the high cost of production [147,155,156].
Production of microalgae is, on the other hand, already an economical
method for sewage treatment [153]. Decreasing the cost of microalgae
production for different purposes, such as a source of oils, polysaccharides, ne chemicals, etc., has been the subject of many studies since
the 1940s [157]. During the 1950s a world-wide interest in novel
sources of protein to feed the growing human population led
researchers to investigate the possibilities of large-scale algal cultivation
systems [152].
Microorganisms, especially bacteria, have been successfully cultured
in large-scale systems such as fermenters for more than half a century
[82]. The basic principles of microalgal cultures are the same as other
microbial cultures with the exception of the light requirement in autotrophic or mixotrophic cultures [158]. For successful microalgal culture,
a suitable species must be selected mainly based on general physical
chemical and biological characteristics together with the growth optimisation of selected species on a suitable medium [159]. On the other
hand, It is well known that science alone is no guarantee that an algal
culture system process or product will be commercially successful
[136,155,160,161]. Commercial success requires the close interaction
between science and business [136]. Once a potential microalga has
been selected through scientic literature surveys and small scale optimisation, then a detailed market survey should be carried out to determine whether there is a market for the product [160]. Computer
modelling of the economies of the process is a powerful tool for decision
making and because of the high costs involved in culturing algae [162],
it allows an estimate of whether the process is economical. Although
large-scale microalgal culture has now been in existence for over
60 years, our experience is still limited to only a few species (i.e.,
Dunaliella, Chlorella, Spirulina) and, even for these, the understanding
of their biology and Ecology is still very limited. The actual cost of
current algal production is generally not revealed by commercial
producers, however, some appreciation of costs can be obtained from
various models which have been published. The total cost for algal
biomass produced as of May 2012 is between 6.20 Aus$ kg 1 and
24 Aus$ kg 1 and normally a lower price can be found in the larger
size plants.
It should be noted that the very low cost presented by some
models (e.g., [163,164]), are from extremely optimistic models and
experience to date shows them to be unrealistic. One of the most successful algae production plants is located at Hutt lagoon, Western
Australia, culturing D. salina for -carotene production. The total
cost of the biomass in this company can be calculated to be between
7.5 Aus$ kg 1 and 13.5 Aus$ kg 1 assuming a 0.09% is the -carotene
concentration of D. salina biomass [165] and a price for -carotene of
approximately 150 Aus$ kg 1 [166].
The cost of conventional fuel (oil) has been growing rapidly
during the last few decades and CO2 emission has also become a
serious environmental issue. Coccolithophorid algae are attractive for
biological CO2 xation and recycling. The oil-rich biomass can be used
as a feedstock for energy production (or possibly as a source of marine
oils) and the CaCO3 plates (coccoliths) produced can be buried to
fossilise the carbon. This is not possible with other non-calcareous
algae. The lipids of these algae also have potential economic value as
marine oils. They are rich in C22:6 3 polyunsaturated fatty acid
[167] and these oils have commercial value in their own rights [168].
The long term reliable small scale outdoor culture of P. carterae CCMP
647 in paddle wheel driven raceway ponds has been already successful
[102]. On the other hand Emiliania huxleyi CCMP 371 outdoor cultivation in open ponds were unsuccessful and it was mainly due to some
limits to the growth of this alga [102,156]. Gephyrocapsa oceanica, P.
carterae and E. huxleyi were grown with reasonably high biomass productivities in plate, airlift and carboy indoor photobioreactors [55].
However due to their shear sensitivity [55] they cannot be grown in
Biocoil photobioreactors. Although the cost of the P. carterae algal

128

N.R. Moheimani et al. / Algal Research 1 (2012) 120133

biomass, as calculated using a simple economic model [56], seems to be


too high to use this alga economically for CO2 bioremediation or biofuel
production at this stage. However, there are several options for reducing this cost and as such, the large-scale culture of the coccolithophorid
alga P. carterae CCMP647 is a serious option for future use for carbon
dioxide bioremediation and possibly as a renewable fuel source.
9. Other commercial applications of coccolithophids
9.1. Immobilisation of CO2 through biomass burial
Burial of the entire wet biomass may be an option for in CO2
removal. The cost effectiveness of such a process and its economical
feasibility would be highly related to value of the carbon credits
obtained and government support for CO2 removal.
9.2. Biofuel
Research on the production of energy from renewable sources,
such as hydrocarbon production by pyrolysis of biomass, has recently
received much recent interest [169,170]. Most renewable sources
used in the pyrolysis of biomass have been higher plants rather
than microalgae, although the latter are the main primary producers
in the oceans [150]. Whereas the depletion of fossil fuels such as
coal and oil lead to the over-emission of CO2, renewable biomass
does not increase atmospheric CO2 concentration and such materials
are now being considered as an important energy resources for the
future [170]. Ginzburg [171] explored the possibility of using algae
such as Dunaliella salina for oil production by biomass pyrolysis and
producing oil 20% less expensive than fossil oil. However Ginsburg's
[171] calculation was based on cells growing with a growth rate of
8.36 d 1 which is not achievable in any type of outdoor cultivation
system. Wu and colleagues [172] also achieved high methane production by high temperature pyrolysis of E. huxleyi. Unfortunately all
reports and research conducted are based on the small scale.
One of the most important factors limiting the growth and improvement of microalgal biotechnology is that there is little prospect for any
alternative designs for microalgae production systems that would be
able to meet the requirements of microalgae production for fuels
[23,155,170]. This is particularly true of closed photobioreactors which
cannot compete with open systems due to high construction and operating costs. The costs of even the simplest closed photobioreactors
would likely be well above what is affordable for the fuel production
process [173,174].
Applying large-scale pond systems for the production of liquid
fuel (biofuel) has been supported by US Department of Energy
(DoE) during the last three decades [175]. The main focus of this project was the production of biodiesel from high-lipid content algae
grown in ponds, utilising waste CO2 from coal red power plants.
While, this project did not result in a large scale production of
microalgae for biofuel production mainly due to a) the sudden change
of the oil economy in the late 70s, b) the high cost of production and
c) low temperature at study site [170]. However it is to be mentioned
that this study provided very useful information on long term growth
of several potential strains of microalgae [170].
The high lipid content and the CaCO3 producing ability of
coccolithophorids could be a potential advantage for large-scale cultivation of such algae for liquid fuels [102,156]. Due to the fact that
conventional fuel resources will nish sooner or later, there is a bright
future for a biofuel production especially using high lipid content
microalgae. In the process of biofuel from biomass production and
irrespective of source of biomass and the conversion methodologies,
the xed C in the biomass will alternatively be returned to the
carbon cycle when being burnt as a fuel. While, every mole of C
xed by any photosynthetic organism, which can be converted to biofuel including but not limited to microalgae, will reduce our dependency

to fossil fuel, this recycled C will stay in the carbon cycle and cannot be
considered as a sequestered C. On the other hand, the main advantage
of coccolithophorid algae compared to any other microalgae is their
capability to x C in both organic and inorganic form as CaCO3
which remains out of carbon cycle. So, coccolithophorid algae culture
can be considered as a potential source for C bioremediation [56,102].
The main question is can coccolithophorids be cultured reliably and
for long period of time under real world conditions? Moheimani and
Borowitzka in a series of studies indicated that P. carterae can be cultured reliably under outdoor conditions in paddle wheel driven raceway
ponds with overall biomass areal productivity of ~19.5 g m2 d1 in
Perth, Western Australia [100,102,156]. This is almost equivalent to a
total biomass productivity of 60 t ha1 y1 with 21.9 t of oil ha1 y1
of and 5.5 t CaCO3 ha1 y1 [100,102,156]. This amount of calcium
carbonate removes 2.6 t of CO2 ha 1 y1 and 0.7 t of Cha1 y1.
According to Diesendorf [176] Annual CO2 emissions from coal red
power stations in 2000 was 186 Mt. Therefore, to be able to bioremediate 0.1% of Australia's annual CO2 emissions to CaCO3 using
culture of P. carterae, there will be a need to over 700 km2 of pond
area. It must be noted that such a plant can also produce 1.6 Mt of
bio-oil, which can be converted to biofuel [177], and 2.3 Mt of high
protein animal feedstock. This potential produced biodiesel can alternatively replace up to 7.2% of annual Australian diesel consumption
of 19,044 ML y1 [178]. It is the authors opinion that a possible
coccolithophorid plant located close to a coal red power plant can not
only minimise the cost but also can reduce the CO2 emission by direct
use of ue gas. The economies of such a system will remain unknown.
Neither E. huxleyi nor G. oceanica could be grown reliably in paddle
wheel driven raceway ponds [102]. As noted earlier, the global importance of E. huxleyi is widely known [36,57,64]. The estimated total
global calcite productivity has been between 0.63 and 1.2 Gt of
calcite y 1 [179]. While, the relative contributions of coccolithophores,
foraminifera, and other calcied organisms remain relatively poorly
known [180], it has been reported that a single large E. huxleyi bloom
can produce up to 7.2 104 tonnes of calcite [68] which is equate to
3.2 104 tonnes of sequestered CO2. As a matter of fact, from oceanographic studies, it has been concluded that coccolithophorids may act
as a net carbon sink through rapid sinking of dead cells to the ocean
oor and burial of the organic and inorganic matter in the ocean
sediments [181]. It is also noted that blooms of E. huxleyi can result in
a net increase of CO2 [127]. This is based on E. huxleyi calcication/
photosynthesis mechanism (see Fig. 5) which can result in lowering
overall pH of the medium [100]. It has been noted that culture of
P. carterae does not follow the same pattern as E. huxleyi. Pleurochrysis
carterae increases the medium pH (up to pH11) [100] which means
that theoretically, in a large scale culture of P. carterae there will be
no net increase of the pH meaning that there will also be no increase
in the net CO2. Based on the current results, it seems that the main reason that P. carterae can grow to pH as high as pH 11 is that this alga has
a very effective CCM mechanism [100,102]. The main reason for the
calcication differences between P. carterae and E. huxleyi is still unknown; however E. huxleyi calcication to photosynthesis (C:P) ratio
is almost 20 times more than P. carterae ([100]. This can explain
why E. huxleyi produce more CO2 and reduce pH in the process of
calcication.

9.3. Biomass coring


The direct co-ring of algal biomass (e.g., coccolithophorids) with a
high lipid and hydrocarbon content with coal has been investigated in
many studies [23,182]. The main advantage of algal co-ring is to
reduce the process costs by increasing efciency of the process. Therefore, in countries such as Australia, where coal plays an important role
for power production, co-ring coal together with microalgae with
high lipid content can increase the efciency.

N.R. Moheimani et al. / Algal Research 1 (2012) 120133

9.4. The application of cell lipids and coccolithophorid algae as pharmaceutical or nutraceuticals
The world population is approximately 7 billion and it is expected to
increase rapidly, so there is need for more nutrients (e.g., protein,
vitamin, lipid, etc.) than can be provided by conventional means.
Photosynthetic aquatic biomass is one of the few resources awaiting
general exploitation [183]. For example, Spirulina by virtue of its high
available protein and vitamin content [184] is considered a favoured
potential source of non-conventional food resources [183]. Over the
last few decades, there has been a large increase in research into alternative sources of fats and oils, particularly the potential commercial
production of microbial lipids and fatty acids [185187]. The bulk of
fats and oils for human consumption and industrial uses are presently
derived from plant sources, and there are still requirements for improved quality and yield, as well as novel products with specialised
chemical composition [188]. The main commercial sources of industrial
oils and fatty acid content are coconut, oil palm and tallow (saturated
fatty acids), olive and canola (mono unsaturated), soybean, castor, linseed and sunower oil (poly-unsaturated) [185].
Oleaginous microorganisms, which may be dened as those which
accumulate greater than 25% of their biomass as oil or fat, have
known to exist for some time [188]. The best microorganisms, capable
of commercial application in lipid production are microalgae, yeasts
and fungi [189191]. High value lipid products for the medical and
pharmaceutical industry are also one of the potential commercial
areas of mass algal production [186]. Currently, the interest in microbial lipids has shifted to polyunsaturated fatty acids, particularly the
polyunsaturated fatty acids (PUFA), docosahexaenoic acid 22:6
(DHA), gamma-linoleic acid (GLA), and arachidonic acid 20:4 (AA),
because of the important role that polyunsaturated fatty acids play
in human nutrition and health [192]. PUFAs, naturally occurring in
marine food chains, are implicated as key therapeutic agents in
reducing the incidence of arthrosclerosis and cardiovascular disease
in humans [193]. Gamma-linoleic acid and arachidonic acid have
strong EPA activity and are being studied for their benecial
treatment of pre-menstrual syndrome, atopic eczema, rheumatoid
arthritis, multiple sclerosis and diabetes [193]. Human populations
whose diets are rich in marine oils derived from sh have been shown
to have reduced levels of serum triacylglycerides and high HDL to
LDL ratios, correlating with a low incidence of cardiovascular disease
[192]. Microbial productions of these PUFAs have the potential
for a larger market beyond traditional sources from sh oil capsules
[194].
Studies of some coccolithophorids have indicated a high lipid
content of up to 40%60% of total biomass [79,195199]. Conte and colleagues [200] divided coccolithophorid algae into two groups based on
their fatty acid variations: (1) a group with high 14:0/16:0 and 16:0/
16:1 ratios, very high 18:43, 18:53, and 22:63 concentrations
(e.g., E. huxleyi); and (2) a group, compromising the Coccolithaceae
and Pleurochrysidaceae, with low 14:0/16:0 and high 16:0/16:0 and
high 18:43 concentrations. Haptophyte algae generally have quite
simple sterol proles, with a single sterol often compromising more
than 75% of the total [200]. Ghosh et al. [201] found high concentrations of the valuable sterol 24-methylcholesta-5,22-dienol, which is
used for bivalve food, in P. carterae. Among the neutral lipids of
coccolithophorid algae, long chain alkenones and alkenoates have
also attracted considerable attention since they occur widely in marine
lacustrine sediments where they are used as biological markers for
inputs of Haptophyte algae [202]. Nevenzel [192] reported that main
hydrocarbons of E. huxleyi are straight-chain polyenes (31:2 and
37:3). Moreover, the coccolithophorid alga P. carterae has been classied to be safe as a supplement in human food in Japan [203]. The
lyophilised, calcium-rich P. carterae, cells have already been used for
human health food as a calcium supplement in Japan for several
years [204]. Miyamoto et al. [205] have also found a considerably

129

high vitamin B12 concentration (0.0013% of dry cell weight) in


lyophilised P. carterae cells.
9.5. Applications of calcium carbonate
Algal calcium carbonate, including coccoliths has become of interest
for possible use in various applications of biomedical science including:
(1) the construction of articial bone in humans [206], (2) complement activation enhancement [207], (3) articial dental root construction [208], (4) scaffolding supports in tissue engineering [209], and
(5) biomedical implants [210]. They have also been targeted by material
scientists having potential signicance as lightweight ceramics, catalyst
support, and robust membranes for high temperature separation technology [210].
10. Conclusion
There are several challenges awaiting us over the next few generations, increase in atmospheric CO2, resulting in possible climate change
and ocean acidication, increased population numbers as well as an
increase demand for energy and an ever reducing amount of freshwater.
While large scale microalgae and coccolithophids culture may not be the
answer to all these issues, they can possibly play be a signicant role in
alleviating our reliance on fossil fuels, and providing a sustainable and secure energy source, as well as providing additional nutrition to the ever
increasing population.
Acknowledgements
This study was funded by an Australian Research Council SPIRT
grant to Michael A Borowitzka and by Rio Tinto.
References
[1] H.-H. Rogner, D. Zhou, R. Bradley, P. Crabb, O. Edenhofer, B. Hare, et al.,
Mitigation. Contribution of Working Group III to the Fourth Assessment Report
of the Intergovernmental Panel on Climate Change, In: O.R.D.B. Metz, P.R.
Bosch, R. Dave, L.A. Meyer (Eds.), Introduction. In Climate Change 2007,
Cambridge University Press, Cambridge, United Kingdom and New York, NY,
USA, 2007, p. 116.
[2] H.J. Herzog, E.M. Drake, Carbon dioxide recovery and disposal from large energy
systems, Annual Review of Energy and the Environment 21 (1996) 145166.
[3] E. Hughes, J. Benemann, Biological fossil CO2 mitigation, Energy Conversion and
Management 38 (1997) S467S473.
[4] P.M. Vitousek, Beyond global warming: ecology and global change, Ecology 75
(1994) 18611876.
[5] IPCC, Climate change 2007: synthesis report. Contribution of Working Groups I, II and
III to the 4th Assessment, In: L.K. Kim JM, K. Shin, J.H. Kang (Eds.), Report of the
Intergovernmental Panel on Climate Change., Cambridge, 2007.
[6] T.R. Society, Ocean acidication due to increasing atmospheric carbon dioxide,
The Royal Society, London, 2005.
[7] K. Schneider, J. Erez, The effect of carbonate chemistry on calcication and
photosynthesis in the hermatypic coral Acropora eurystoma, Limnology and
Oceanography 51 (2006) 12841293.
[8] W.C. Turkenburg, Sustainable development, climate change, and carbon dioxide
removal (CDR), Energy Conversion and Management 38 (1997) S3S12.
[9] R.L. Kane, D.E. Klein, United States strategy for mitigating global climate change,
Energy Conversion and Management 38 (1997) S13S18.
[10] J. Benemann, CO2 mitigation with microalgae systems, Energy Conversion and
Management 38 (1997) 475479.
[11] A. Meisen, X. Shuai, Research and development issues in CO2 capture, Energy
Conversion and Management 38 (1997) S37S42.
[12] D.F. Spencer, Ocean systems for managing the global carbon cycle, Energy
Conversion and Management 38 (1997) S265S271.
[13] C. Marchetti, On geoengineering and the CO2 problem, Climatic Change 1 (1977)
5968.
[14] T. Ohsumi, Introduction: what is the ocean sequestration of carbon dioxide?
Journal of Oceanography 60 (2004) 693694.
[15] R. Hase, H. Oikawa, C. Sasso, M. Morita, Y. Watanabe, Photosynthetic production
of microalgal biomass in a raceway system under greenhouse conditions in
Sendi city, Journal of Bioscience and Bioengineering 89 (2000) 157163.
[16] K.T. Klasson, B.H. Davison, Estimation of carbon credits in the carbon dioxide
sequestration activities, In: K.T. Klasson (Ed.), First National Conference on
Carbon Sequestration, U.S Department Energy, Washington, DC, 2001, pp. 117.

130

N.R. Moheimani et al. / Algal Research 1 (2012) 120133

[17] A. Richmond, N. Zou, Efcient utilisation of high photon irradiance for mass
production of photoautotrophic micro-organisms, Journal of Applied Phycology
11 (1999) 123127.
[18] M.R. Tredici, R. Materassi, From open ponds to vertical alveolar panels: the
Italian experience in the development of reactors for the mass cultivation
of phototrophic microorganisms, Journal of Applied Phycology 4 (1992)
221231.
[19] M.A. Borowitzka, N. Moheimani, Sustainable biofuels from algae, Mitigation and
Adaptation Strategies for Global Change (2010), http://dx.doi.org/10.1007/
s11027-010-9271-9.
[20] S. Fon Sing, A. Isdepsky, M. Borowitzka, N. Moheimani, Production of biofuels
from microalgae, Mitigation and Adaptation Strategies for Global Change
(2011) 126, http://dx.doi.org/10.1007/s11027-011-9294-x.
[21] A. Anderson, Cost-effectiveness in addressing the "CO2 problem," with special
reference to the investments of the global environment facility, Annual Review
of Energy and the Environment 19 (1994) 423455.
[22] K.S. Lackner, A guide to CO2 sequestration, Science 300 (2003) 16771678.
[23] K.L. Kadam, Power plant gas as a source of CO2 for microalgae cultivation
economic impact of different process options, Energy Conversion and Management
38 (1997) S505S510.
[24] J.C. Green, R.W. Jordan, Systematic history and taxonomy, In: J.C. Green, B.S.C.
Leadbeater (Eds.), The Haptophyte algae, Clarendon Press, Oxford, 1994,
pp. 123.
[25] W.G. Siesser, Historical background of coccolithophore studies, In: A. Winter,
W.G. Siesser (Eds.), Coccolithophores, Cambridge University Press, Cambridge,
1994, pp. 112.
[26] J.R. Young, J.A. Bergen, P.R. Bowen, J.a. Burnett, A. Fiorentino, R.W. Jordan, et al.,
Guidelines for coccolith and calcareous nannofossil terminology, Palaentology
40 (1997) 875912.
[27] L. Cros, A. Kleijne, A. Zeltner, C. Billard, J.R. Young, New example of
holococcolithheterococcolith combination coccospheres and their implications
for coccolithophorid biology, Marine Micropaleontology 39 (1999) 134.
[28] M. Geisen, C. Billard, A.T.C. Broerse, L. Cros, I. Probert, J.R. Young, Life-cycle
association involving pairs of holococcolithophorid species: intraspecic variation
or cryptic specication? European Journal of Phycology 37 (2002) 531550.
[29] J.R. Young, S.A. Davis, P.R. Bown, S. Mann, Coccolith ultrastructure and biomineralization, Journal of Structural Biology 126 (1999) 195215.
[30] C. Billard, Life cycles, In: J.C. Green, B.S.C. Leadbeater (Eds.), The Haptophyte
Algae, Clarendon Press, Oxford, 1994, pp. 167186.
[31] J.R. Young, P. Ziveri, Calculation of coccolith volume and its use in calibration of
carbonate ux estimates, Deep Sea Research Part II: Topical Studies in Oceanography 47 (2000) 16791700.
[32] A. Houdan, C. Billard, D. Marie, F. Not, A.G. Saez, J.R. Young, et al.,
Holococcolithophoreheterococcolithophore (Haptophyta) life cycles: ow
cytometry analysis of relative ploidy levels, Systematics and Biodiversity 1
(2004) 453465.
[33] B.S.C. Leadbeater, Observations on the life history of the haptophycean alga
Pleurochrysis-scherffelii with special reference to the micro anatomy of the
different types of motile cells, Annals of Botany (London) 35 (1971) 429439.
[34] J.D. Rowson, B.S.C. Leadbeater, Calcium carbonate deposition int he motile
(Crystallolithus) phase of Coccolithus pelagicus (Prymnesiophyceae), British
Phycological Journal 21 (1986) 359370.
[35] R.N. Pienaar, Ultrastructure and calcication of coccolithophores, In: A. Winter,
W.G. Siesser (Eds.), Coccolithophores, Cambridge University Press, Cambridge,
1994, pp. 1337.
[36] R.W. Jordan, Haptophyta, els, Yamagata University, Yamagata, Japan, 2012.
[37] J.C. Green, P.A. Course, G.A. Tarran, The life-cycle of Emiliania huxleyi: a brief
review and a study of relative ploidy levels analysed by ow cytometry, Journal
of Marine Systems 9 (1996) 3344.
[38] E.K. Hawkins, J.J. Lee, D.K. Fimiarz, Colony formation and sexual morphogenesis
in the coccolithophore Pleurochrysis sp. (Haptophyta) 1, Journal of Phycology 47
(2011) 13441349.
[39] P. Gayral, J. Frensel, Description sexuality and developmental cycle of a new
coccolithophorid Prymnesiophyceae Pleurochrysis-pseudoroscoffensis new-species,
Protistologica 19 (1983) 245262.
[40] M. Nel, M. Kawachi, I. Inouye, Induced dimorphic life cycle of a coccolithophorid,
Calyptrosphaera sphaeroidea (Prymnesiophyceae, Haptophyta), Journal of Phycology 40 (2004) 112129.
[41] D.G. Rayns, Alternation of generations in a coccolithophorid, Cricosphaera carterae
(Braarud & Fagerl.) Braarud, Journal of the Marine Biological Association of the UK
42 (1962) 481484.
[42] B.S.C. Leadbeater, Preliminary observations on differences of scale morphology
at various stages in the life cycle of Apistonema syracosphaer sensu von Stosch,
British Phycological Journal 5 (1970) 5769.
[43] P. Gayral, J. Frensel, New observations on 2 marine coccolithophoraceae
Cricosphaera roscoffensis new combination and Hymenomonas globosa new
combination, Phycologia 15 (1976) 339356.
[44] J. Fresnel, C. Billard, Pleurochrysis placolithoides sp. nov. (Prymnesiophyceae), a
new marine coccolithophorid with remarks on the status of criocolith-bearing
species, British Phycological Journal 26 (1991) 6780.
[45] I. Manton, Further observations on the ne structure of Chrysochromulina chiton
with special reference to the haptonema, peculiar Golgi structure and scale
production, Journal of Cell Science 2 (1967) 265272.
[46] M. Kawachi, I. Inouye, Functional roles of the haptonema and the spine scales in
the feeding process of Chrysochromnlina spinifera(Fournier) Pienaar et Norris
(Haptophyta, Prymnesiophyta), Phycologia 34 (1995) 193200.

[47] E. Paasche, Biology and physiology of coccolithophorids, Annual Review of


Microbiology 22 (1968) 7186.
[48] J. Fresnel, I. Probert, The ultrastructure and life cycle of the coastal
coccolithophorid Ochrosphaera neapolitana (Prymnesiophyceae), European
Journal of Phycology 40 (2005) 105122.
[49] P.J. Hansen, M. Hjorth, Growth and grazing responses of Chrysochromulina
ericina (Prymnesiophyceae): the role of irradiance, prey concentration and pH,
Marine Biology 141 (2002) 975983.
[50] I. Inouye, R.N. Pienaar, Ultrastructure of the agellar apparatus in Pleurochrysis
(Class Prymnesiophyceae), Protoplasm 125 (1985) 2435.
[51] S. Sym, M. Kawachi, Ultrastructure of Calyptrosphaera radiata, sp nov
(Prymnesiophyceae, Haptophyta), European Journal of Phycology 35 (2000)
283293.
[52] M.A. Borowitzka, Algal calcication, Oceanography and Marine Biology. Annual
Review 15 (1977) 189223.
[53] C.S. Sikes, R.D. Roer, K.M. Wilber, Photosynthesis and coccolith formation:
Inorganic carbon sources and net inorganic reaction of deposition, Limnology
and Oceanography 25 (1980) 248261.
[54] L.E. Brand, Physiological ecology of marine coccolithophores, In: A. Winter, W.G.
Siesser (Eds.), Coccolithophores, Cambridge University Press, Cambridge, 1994,
pp. 3949.
[55] N.R. Moheimani, A. Isdepsky, J. Lisec, E. Raes, M.A. Borowitzka, Coccolithophorid
algae culture in closed photobioreactors, Biotechnology and Bioengineering 108
(2011) 20782087.
[56] N.R. Moheimani. The culture of coccolithophorid algae for carbon dioxide
bioremediation [PhD Thesis]: Murdoch University, 2005.
[57] E. Paasche, A review of the coccolithophorid Emiliania huxleyi (Prymnesiophyceae),
with particular reference to growth, coccolith formation, and calcicationphotosynthesis interactions, Phycologia 40 (2002) 503529.
[58] E. Paasche, S. Brubak, S. Skattebol, J.R. Young, J.C. Green, Growth and calcication
in the coccolithophorid Emiliania huxleyi (Haptophyceae) at low salinities,
Phycologia 35 (1996) 394403.
[59] M.V. Nielsen, Photosynthetic characteristic of the coccolithophorid Emiliania
huxleyi (Prymnesiophyceae) exposed to elevated concentrations of dissolved
inorganic carbon, Journal of Phycology 31 (1995) 715719.
[60] J.R. Johansen, G.J. Doucette, W.R. Barclay, J.D. Bull, The morphology and ecology
of Pleurochrysis carterae var dentata var nov. (Prymnesiophyceae), a new
coccolithophorid from an inland saline pond in New Mexico, USA, Phycologia
27 (1988) 7888.
[61] E.W. de Vrind-de Jong, A. Borman, P. Thierry, P. Westbroek, Gruter, J.P. Kamerling,
Calcication in the coccolithophorids Emiliania huxleyi and Pleurochrysis carterae
II. Biochemical aspects, In: B.S.C. Leadbeater (Ed.), Bioremediation in Lower Plants
and Animals, Clarendon Press, Oxford, 1986, pp. 205217.
[62] E.W. de Vrind, P.R. de Jong, P.R. van Emburg, J.P.M. de Vrind, Mechanisms of calcication: Emiliania huxleyi as a model system, In: J.C. Green, B.S.C. Leadbeater (Eds.), The
Haptophyte Algae, Clarendon Press, Oxford, 1994, pp. 149166.
[63] P. van der Wal, L. de Jong, P. Westbroek, W.C. de Bruijn, Calcication in the
coccolithophorid alga Hymnemonas carterae, In: P. van der Wal (Ed.), Calcication
in Two species of coccolithophorid algae, vol. 1, Geologisch Institute, Amsterdam,
1984, pp. 7380.
[64] M.E. Marsh, Regulation of CaCO3 formation in coccolithophores, Comparative
Biochemistry and Physiology. B: Comparative Biochemistry 136 (2003) 743754.
[65] M. Okazaki, M. Fujii, Y. Usuda, K. Furuya, Soluble Ca2+-activated ATPase and its
possible role in calcication of the coccolithophorid Cricosphaera roscoffensis
var. haptonemofera (Haptophyta) (studies on the calcium carbonate deposition
of algae V), Botanica Marina 27 (1984) 363369.
[66] C. Linschooten, J.D.L. van Bleijwijk, P.r. van Emburg, J.p.m. de Vrind, E.S. Kempers,
P. Westbroek, et al., Role of the lightdark cycle and medium composition on the
production of coccoliths by Emiliania huxleyi (Haptophyceae), Journal of Phycology
27 (1991) 8286.
[67] S.W. Jeffrey, M.B. Allen, Pigments, growth and photosynthesis in cultures of two
chrysomonads, Coccolithus huxleyi and a Hymenomonassp, Journal of General
Microbiology 36 (1964) 277288.
[68] P.M. Holligan, M. Viollier, D.S. Harbour, P. Camus, M. Champagne-Philippe,
Satellite and ship studies of coccolithophore production along a continental
shelf edge, Nature 304 (1983) 339342.
[69] S.I. Blackburn, G. Cresswell, A coccolithophorid bloom in Jervis Bay, Australia,
Australian Journal of Marine & Freshwater Research 44 (1993) 253260.
[70] W.M. Balch, K.A. Kilpatrick, C.C. Trees, The 1991 coccolithophore bloom in the
central North Atlantic, 1. Optical properties and factors affecting their distribution,
Limnology and Oceanography 41 (1996) 1831669.
[71] K.M. Reifel, M.P. McCoy, M.A. Tiffany, T.E. Rocke, C.C. Trees, S.B. Barlow, et al.,
Pleurochrysis pseudoroscoffensis (Prymnesiophyceae) blooms on the surface of
the Salton Sea, California, Hydrobiologia 466 (2001) 177185.
[72] E.M. Hulburt, The diversity of phytoplanktonic populations in oceanic, coastal,
and estuarine regions, Journal of Marine Research 21 (1963) 8193.
[73] A. McIntyre, Coccoliths as paleoclimatic indicators of Pleistocene glaciation,
Science 158 (1967) 13141317.
[74] E.T. Buitenhuis, H.J.W. de Baar, M.J.W. Veldhuis, Photosynthesis and calcication
by Emiliania huxleyi (Prymnesiophyceae) as a function of inorganic carbon
species, Journal of Phycology 35 (1999) 949959.
[75] D.A. Purdie, M.S. Finch, Impact of a coccolithophorid bloom on dissolved carbon
dioxide in seawater enclosures in a Norwegian Fjord, Sarsia 79 (1994) 24379.
[76] A. Winter, R.W. Jordan, P.H. Roth, Biogeography of living coccolithophores in
ocean waters, In: A. Winter, W.G. Siesser (Eds.), Coccolithophores, Cambridge
University Press, Cambridge, 1994.

N.R. Moheimani et al. / Algal Research 1 (2012) 120133


[77] E. Miyamoto, F. Watanabe, H. Takenaka, Y. Nakano, Uptake and physiological
function of vitamin B-12 in a photosynthetic unicellular coccolithophorid alga,
Pleurochrysis carterae, Bioscience, Biotechnology, and Biochemistry 66 (2002)
195198.
[78] M. Parke, I. Adams, The motile (Crystallolithus hyalinus Gaarder and Markali) and
non-phases in the life history of Coccolithus pelagicus (Wallich) Schiller,
Journal of the Marine Biological Association of the United Kingdom 39 (1960)
263274.
[79] P. Corstjens, Y. Araki, E.L. Gonzalez, A coccolithophorid calcifying vesicle with a
vacuolar-type ATPase proton pump: cloning and immunolocalization of the V-0
subunit c(1), Journal of Phycology 37 (2001) 7178.
[80] N.A. Nimer, J.M. Merret, Calcication rate in Emiliania huxleyi Lohman in
response to light, nitrate and availability of inorganic carbon, New Phytologist
123 (1993) 673677.
[81] E. Paasche, Roles of nitrogen and phosphorous in coccolith formation in
Emiliania huxleyi, European Journal of Phycology 33 (1998) 3342.
[82] J.J. Fritz, Carbon xation and coccolith detachment in the coccolithophore
Emiliania huxleyi in nitrate-limited cyclostats, Marine Biology 133 (1999)
509518.
[83] M.Y. Cortes, J. Bollmann, H.R. Thierstein, Coccolithophore ecology at the HOT
station ALOHA, Hawaii, Deep Sea Research Part II: Topical Studies in Oceanography
48 (2001) 19571981.
[84] A.T. Haidar, H.R. Thierstein, Coccolithophore dynamics off Bermuda (N. Atlantic),
Deep Sea Research Part II: Topical Studies in Oceanography 48 (2001) 19251956.
[85] Y. Shiraiawa, Physiological regulation of carbon xation in the photosynthesis
and calcication of coccolithophorids, Comparative Biochemistry and Physiology
136 (2003) 775783.
[86] L.E. Brand, W.G. Sunda, R.R.L. Guillard, Limitation of marine phytoplankton reproductive rates by zinc, manganese and iron, Limnology and Oceanography
28 (1983) 11821198.
[87] L.E. Brand, Minimum iron requirements of marine plhytoplankton and the
implications for the biogeochemical control of new production, Limnology and
Oceanography 36 (1991).
[88] T. McConnaughey, Acid secreation, calcication, and photosynhtetic carbon
concentrating mechanisms, Canadian Journal of Botany-Revue Canadienne de
Botanique 76 (1998) 11191126.
[89] E.G. Jorgensen, E. Steemann Nielsen, Effect of ltrates from cultures of unicellular
algae on the growth of Staphylococcus aureus, Physiologia Plantarum 14 (1961)
896908.
[90] A. Houdan, A. Bonnard, J. Fresnel, C. Fouchard, C. Billard, I. Probert, Toxicity of
coastal coccolithophores (Prymnesiophyceae, Haptophyta), Journal of Plankton
Research 26 (2004) 875883.
[91] K. Reifel, M. McCoy, M. Tiffany, T. Rocke, C. Trees, S. Barlow, et al., Pleurochrysis
pseudoroscoffensis (Prymnesiophyceae) blooms on the surface of the Salton
Sea, California, Hydrobiologia 466 (2001) 177185.
[92] H.L. Ehrlich, Geomicrobiology its signicance for geology, Earth-Science Reviews
45 (1998) 4560.
[93] J. Guttusso, R.W. Buddemeier, Calcication and CO2, Nature 407 (2000) 311312.
[94] W.W. Faber, H.R. Preisig, Calcied structures and calcication in protists,
Protoplasma 181 (1994) 78105.
[95] M.A. Borowitzka, Mechanisms in algal calcication, Progress of Phycological
Research 1 (1982) 137177.
[96] C. Brownlee, N.A. Nimer, L.F. Dong, J.M. Merret, Cellular regulation during calcication
in Emiliania huxleyi, In: J.C. Green, B.S.C. Leadbeater (Eds.), The Haptophyte Algae,
Clarendon Press, Oxford, 1994, pp. 133148.
[97] W. Stumm, J.J. Morgan, Aquatic chemistry: chemical equilibria and rates in
natural waters, John Wiley and Sons, Inc., 1996.
[98] J.A. Raven, Putting C in phycology, European Journal of Phycology 32 (1997)
319333.
[99] S.-y. Fukuda, I. Suzuki, T. Hama, Y. Shiraiwa, Compensatory response of the
unicellular-calcifying alga Emiliania huxleyi (Coccolithophoridales, Haptophyta)
to ocean acidication, Journal of Oceanography 67 (2011) 1725.
[100] N. Moheimani, M. Borowitzka, Increased CO2; and the effect of pH on growth
and calcication of Pleurochrysis carterae and Emiliania huxleyi (Haptophyta)
in semicontinuous cultures, Applied Microbiology and Biotechnology 90
(2011) 13991407.
[101] Y. Shiraiwa, A. Goyal, N.E. Tolbert, Alkalization of the medium by unicellular
green algae during uptake dissolved inorganic carbon, Plant & Cell Physiology
34 (1993) 649657.
[102] N. Moheimani, M. Borowitzka, The long-term culture of the coccolithophore
Pleurochrysis carterae (Haptophyta) in outdoor raceway ponds, Journal of
Applied Phycology 18 (2006) 703712.
[103] J.U. Grobbelaar, Algal nutrition, mineral nutrition, In: A. Richmond (Ed.),
Microalgal Culture, Blackwell Science Ltd., Oxford, 2004, pp. 97115.
[104] P.g. Brewer, J.C. Goldman, Alkalinity changes generated by phytoplankton
growth, Limnology and Oceanography 21 (1976) 108117.
[105] E.A. Laws, P.a. Thompson, B.N. Popp, R.R. Bidigare, Sources of inorganic carbon
for marine microalgal photosynthesis: a reassessment of S13C data from batch
culture studies of Thalassiosira pseudonana and Emiliania huxleyi, Limnology
and Oceanography 43 (1988) 136142.
[106] J.V. Moroney, Carbon concentrating mechanisms in aquatic photosynthetic
organisms: a report on CCM 2001, Journal of Phycology 37 (2001) 928931.
[107] J. Beardall, A. Johnston, J. Raven, Environmental regulation of CO2-concentrating
mechanisms in microalgae, Canadian Journal of Botany 76 (1998) 10101017.
[108] J.A. Raven, P.G. Falkowski, Oceanic sinks for atmospheric CO2, Plant, Cell &
Environment 22 (1999) 741755.

131

[109] K. Sekino, Y. Shiraiawa, Accumulation and utilization of dissolved inorganic carbon


by a marine unicellular coccolithophorid, E. huxleyi, Plant & Cell Physiology 35
(1994) 353361.
[110] E. Paasche, S. Brubak, Enhanced calcication in the coccolithophorid Emiliania
huxleyi (Haptophyceae) under phosphorus limitation, Phycologia 33 (1994)
324330.
[111] L.F. Dong, N.A. Nimer, E. Okus, M.J. Merrett, Dissolved inorganic carbon utilization
in relation to calcite production in Emiliania huxleyi (Lohmann) Kamptner, New
Phytologist 123 (1993) 679684.
[112] M. Frankignoulle, C. Canon, J.P. Gattuso, Marine calcication as a source of carbon
dioxide: positive feedback of increasing CO2, Limnology and Oceanography 39
(1994) 458462.
[113] M. Badger, The roles of carbonic anhydrases in photosynthetic CO2 concentrating
mechanisms, Photosynthesis Research 77 (2003) 8394.
[114] T. Anning, N. Nimer, M.J. Merrett, C. Brownlee, Costs and benets of calcication
in coccolithophorids, Journal of Marine Systems 9 (1996) 4556.
[115] J.A. Raven, Inorganic carbon concentrating mechanisms in relation to the biology
of algae, Photosynthesis Research V77 (2003) 155171.
[116] A. Kaplan, L. Reinold, CO2 Concentrating mechanism in photosynthetic microorganisms, Annual Review of Plant Physiology and Plant Molecular Biology 50
(1999) 539570.
[117] I.E. Huertas, S. Bhatti, B. Colman, Inorganic carbon acquisition in two species of
marine prymnesiophytes, European Journal of Phycology 38 (2003) 181189.
[118] C.S. Sikes, A.P. Wheeler, Carbonic anhydrase and carbon xation in coccolithophorids,
Journal of Phycology 18 (1982) 423426.
[119] N.A. Nimer, Q. Guan, M.J. Merrett, Extra- and intra- cellular carbonic anhydrase
in relation to culture age in a high-calcifying strain of Emiliania huxleyi, New
Phytologist 126 (1994) 601607.
[120] N.A. Nimer, J.M. Merret, Inorganic carbon transport in relation to culture age and
inorganic carbon concentration in a high-calcifying strain of Emiliania huxleyi
(Prymnesiophyceae), Journal of Phycology 32 (1996) 813818.
[121] L. Herfort, B. Thake, J. Roberts, Acquisition and use of bicarbonate by Emiliania
huxleyi, New Phytologist 156 (2002) 427436.
[122] A.A. Israel, E.L. Gonzales, Photosynthesis and inorganic carbon utilization in
Pleurochrysis sp. (Haptophyta), a coccolithophorid alga, Marine Ecology Progress
Series 137 (1996) 243250.
[123] E. Paasche, A tracer study of the inorganic carbon uptake during coccolith formation
and photosynthesis in the coccolithophorid Coccolithus huxleyi, Physiologia
Plantarum 3 (1964) 182.
[124] Y. Shiraiawa, H. Sugimoto, M. Satoh, Regulation of intracellular calcication and
algal growth by nutrient supply in coccolithophorids, In: I. Kobayashi, H. Ozawa
(Eds.), Proceedings of the 8th international symposium on biomineralization
(biomineralization: formation, diversity, evolution and application), Tokai University Press, Kangawa, 2003, pp. 241246.
[125] L. Berry, A.R. Taylor, U.E. Lucken, K.P. Ryan, C. Brownlee, Calcication and inorganic carbon acquisition in coccolithophores, Functional Plant Biology 29
(2002) 23.
[126] M.A. Borowitzka, Carbonate calcication in algae initiation and control, In: S.
Mann, J. Webb (Eds.), Biomineralization: Chemical and Biochemical Perspective,
VCH Verlag, Weinheim, 1989, pp. 6394.
[127] U. Riebesell, I. Zondervan, B. Rost, P.D. Tortell, R.E. Zeebe, F.M.M. Morel, Reduced
calcication of marine plankton in response to increased atmospheric CO2, Nature 407 (2000) 364367.
[128] D.W. Crawford, D.A. Purdie, Increase of P CO2 during blooms of Emiliania huxleyi:
theorical considerations on the asymmetry between acquisition of HCO3- and
respiration of free CO2, Limnology and Oceanography 42 (1997) 365372.
[129] M.A. Crenshaw, Coccolith formation by two marine coccolithophorids, Duke
University, 1964.
[130] N.A. Nimer, G.K. Dixon, M.J. Merrett, Utilization of inorganic carbon by the
coccolithophorid Emiliania huxleyi (Lohman) Kamptner, New Phytologist 120
(1991) 153158.
[131] J.A. Raven, Nitrogen assimilation and transport in vascular land pklants in
relation tintracellular pH regulation, New Phytologist 76 (1980) 415431.
[132] R.J. Kingsley, W. Norimitsu, Ca-ATPase localization and inhibition in the gorgonian
Leptogorgia virgulata (Lamarck) (Coelenterata: Gorgonacea), Journal of Experimental Marine Biology and Ecology 93 (1985) 157167.
[133] D. Klaveness, Emiliania huxleyi (Lohman) Hay and Mohler. 3. Minral deposition
and the origin of the matrix during coccolith formation, Protistologica 12
(1976) 217224.
[134] M. Okazaki, Some enzymatic properties of Ca2+dependent adenosine
triphosphatase from calcareous marine alga, Serraticardia maxima and its distribution in marine algae, Botanica Marina 20 (1977) 347354.
[135] P. van der Wal, W.C. de Brujin, Calcication in the coccolithophorid alga
Hymnemonas carterae, Environmental Biogeochemistry 35 (1983) 251258.
[136] S.W. Jeffrey, S.W. Wright, Photosynthetic pigments in the Haptophyta, In: Z.
Cohen (Ed.), Chemicals from Microalgae, Taylor and Francis, Philadelphia,
1999, pp. 111132.
[137] L.E. Brand, R.R.L. Guillard, The effects of continuous light and intensity on the
reproduction rates of twenty-two species of marine phytoplankton, Journal of
Experimental Marine Biology and Ecology 50 (1981) 119132.
[138] A.G.J. Buma, T. van Oijen, W. van de Poll, M.J.W. Veldhuis, W.W.C. Gieskes, The
sensitivity of Emiliania huxleyi (Prymnesiophyceae) to ultraviolet-b radiation,
Journal of Phycology 36 (2000) 296303.
[139] W.W.C. Gieskes, A.G.J. Buma, UV damage to plant life in a photobiologically
dynamic environment: the case of marine phytoplankton, Plant Ecology 128
(1997) 1725.

132

N.R. Moheimani et al. / Algal Research 1 (2012) 120133

[140] K.J. Voss, W.M. Balch, K.A. Kilpatrick, Scattering and attenuation properties of
Emiliania huxleyi cells and their detached coccoliths, Limnology and Oceanography
43 (1998) 870876.
[141] J.R. Young, P.R. Brown, J.A. Burnett, Paleontological perspectives, In: J.C. Green,
B.S.C. Leadbeater (Eds.), The Haptophyte Algae, 51 ed., Clarendon Press, Oxford,
UK, 1994, p. 446.
[142] E. Paasche, Adjustment to light and dark rates of coccolith formation,
Physiologia Plantarum 19 (1966) 271278.
[143] K. Sekino, Y. Shiraiawa, Evidence for the involvement of mitochondrial respiration
in calcication in a marine coccolithophorid, Emiliania huxleyi, Plant & Cell
Physiology 37 (1996) 10301033.
[144] W. Balch, J. Fritz, E. Fernandez, Decoupling of calcication and photosynthesis in
the coccolithophore Emiiania huxleyi under steady-state light-limited growth,
Marine Ecology Progress Series 142 (1996) 8797.
[145] P. van der Wal, J.P.M. de Vind, E.W. de Vrind-de Jong, A. Borman, Incompleteness
of the coccosphere as a possible stimulus for coccolith formation in Pleurochrysis
carterae (Prymnesiophyceae), Journal of Phycology 23 (1987) 218221.
[146] J.C. Dodd, Elements of pond design and construction, In: A. Richmond (Ed.), CRC
Handbook of Microalgal Mass Culture, CRC Press, Inc., 1986, pp. 265284.
[147] M. Borowitzka, N. Moheimani, Sustainable biofuels from algae, Mitigation and
Adaptation Strategies for Global Change (2010) 113, http://dx.doi.org/10.1007/
s11027-010-9271-9.
[148] A.W.D. Larkum, Marine primary productivity, In: M.N. Clayton, R.J. King (Eds.),
Marine Botany: An Australasian Perspective, Longman Cheshire, Melbourne,
1981, pp. 370385.
[149] H. Leith, Primary production of the major vegetation units of the world, In: H.
Leith, R.H. Whittacker (Eds.), Primary Productivity of the Biosphere, SpringerVerlag, New York, 1975.
[150] R.J. Geider, E.H. Delucia, P. Falkowski, A. Finzi, P. Grime, J.T.M. Kana, J. La Roche,
S.P. Long, B.A. Osborne, T. Platt, I.C. Prentice, J.A. Raven, W.H. Schlesinger, V.
Smetacek, V. Stuart, S. Sathyendranath, R.B. Thomas, T.C. Vogelmann, P.
Williams, F.I. Woodward, Primary productivity of planet earth: biological determinants and physical constraints in terrestrial and aquatic habitats, Global
Change Biology 7 (2001) 849882.
[151] I. Inouye, M. Chihara, Laboratory culture and taxonomy of Hymenomonas-coronata
and Ochrosphaera-verrucosa class prymnesiophyceae from the Northwest Pacic,
Botanical Magazine Tokyo 93 (1980) 195208.
[152] C.J. Soeder, The scope of microalgae for food and feed, In: G. Shelef, C.J. Soeder
(Eds.), Algae Biomass, Elsevier/North-Holland Biomedical Press, Amsterdam,
1980, pp. 920.
[153] W.J. Oswald, Algal productionproblems, acheivements and potential, In: G.
Shelef, C.J. Soeder (Eds.), Algae Biomass, Elsevier/North-Holand Biomedical
Press, Amsterdam, 1980, pp. 19.
[154] J.R. Benemann, D.M. Tillett, J.C. Weissman, Microalgae biotechnology, Trends in
Biotechnology 5 (1987) 4753.
[155] M.A. Borowitzka, Algal biotechnology products and processesmatching
science and economics, Journal of Applied Phycology 4 (1992) 267279.
[156] N.R. Moheimani, M.A. Borowitzka, Limits to productivity of the alga Pleurochrysis
carterae (Haptophyta) grown in outdoor raceway ponds, Biotechnology and
Bioengineering 96 (2006) 2736.
[157] W. Fulks, K.L. Main, Part I: background review the design and operation of
commercial-scale live feeds production systems, In: W. Fulks, K.L. Main (Eds.),
Rotifer and Microalgae Culture Systems, The Oceanic Institute Makapuu Point,
Honolulu, 1991, pp. 246.
[158] Y.K. Lee, H. Shen, Basic culturing techniques, In: A. Richmond (Ed.), Microalgal
Culture, Blackwell Science Ltd, Oxford, 2004, pp. 4056.
[159] M.A. Borowitzka, Culturing microalgae in outdoor ponds, In: A. Anderson (Ed.),
Algal Culturing Techniques, Academic Press, London, 2005, pp. 205217.
[160] M.A. Borowitzka, Microalgae for aquaculture: opportunities and constraints,
Journal of Applied Phycology 9 (1997) 393401.
[161] A. Sukenik, R.S. Levy, Y. Levy, P.G. Falkowski, Z. Dubinsky, Optimizing algal
biomass production in an outdoor pond: a simulation model, Journal of Applied
Phycology 3 (1991) 191201.
[162] K.E. Apt, P.W. Behrens, Commercial developments in microalgal biotechnology,
Journal of Phycology 35 (1999) 215226.
[163] G. Valderrama, A. Cardenas, A. Markovits, On the economics of Spirulina production
in Chile with details on drag-board mixing in shallow ponds, Hydrobiologia 151
(152) (1987) 7174.
[164] W.R. Barclay, K.L. Terry, N.J. Nagle, J.C. Weissman, R.P. Goebel, Potential of new
strains of marine and inland saline-adapted microalgae for aquaculture, Journal
World of Aquatic Society 18 (1987) 218228.
[165] L.J. Borowitzka, Commercial pigment from algae, In: P.S. Moi, L.Y. Kun, M.A.
Borowitzka, B.A. Whitton (Eds.), The First Asia-Pacic Conference on Algal
Biotechnology, Institute of Advanced Studies, University of Malaya, Kuala
Lumpur, Kuala Lumpur, 1994, pp. 8284.
[166] M.A. Borowitzka, Commercial production of microalgae: ponds, tanks, tubes and
fermenters, Journal of Biotechnology 70 (1999) 313321.
[167] J.K. Volkman, D.J. Smith, G. Eglinton, T.E.V. Forsberg, E.D.S. Corner, Sterol and
fatty acid composition of four marine haptophycean algae, Journal of the Marine
Biological Association of the UK 61 (1981) 509527.
[168] M.A. Borowitzka, L.J. Borowitzka, Dunaliella, In: M.A. Borowitzka, L.J. Borowitzka
(Eds.), Microalgal Biotechnology, Cambridge University Press, Sydney, 1988,
pp. 2758.
[169] J.C. Weissman, R.P. Goebel, Design and analysis of microalgal pond systems for
purpose of producing fuel, In: S.E.R. Institute (Ed.), U.S. Department of Energy,
1987, p. 140.

[170] J. Sheehan, T. Dunahay, J. Benemann, P. Roessler, A look back at the U.S. Department
of Energy's Aquatic Species Program-biodiesel from algae, In: U.S. Department of
Energy's Ofce of Fuels Development, 1998, p. 294.
[171] B. Ginzburg, Liquid fuel (oil) from halophilic algae: a renewable source of
non-polluting energy, Renal Energy 3 (1993) 249252.
[172] Q. Wu, J. Dai, Y. Shiraiawa, G. Sheng, J. Fu, A renewable energy source-hydrocarbon
gases resulting from pyrolysis of the marine nonoplanktonic alga Emiliania huxleyi,
Journal of Applied Phycology 11 (1999) 137142.
[173] A. Richmond, Microalgal biotechnology at the turn of the millennium: a personal
view, Journal of Applied Phycology 12 (2000) 441451.
[174] M.A. Borowitzka, Closed algal photobioreactors: design considerations for
large-scale systems, Journal of Marine Biotechnology 4 (1996) 185191.
[175] J.C. Weisman, J. Noue, Design and operation of an outdoor microalgae test
facility: large-scale system results, Nature Renal Energy Laboratory (1995)
3256.
[176] M. Diesendorf, Australia's Polluting Power: Coal-red Electricity and Its Impact
on Global Warming, In: WWF Australia, Sydney, 2003, p. 12.
[177] K. de Boer, N. Moheimani, M. Borowitzka, P. Bahri, Extraction and conversion
pathways for microalgae to biodiesel: a review focused on energy consumption,
Journal of Applied Phycology (2012) 118, http://dx.doi.org/10.1007/s10811012-9835-z.
[178] RET, Australian Petroleum Statistics, Department of Resources, Energy and Tourism
Canberra, 2012.
[179] IPCC, Contribution of Working Groups I, II and III to the Fourth Assessment
Report of the Intergovernmental Panel on Climate Change, In: R.K. Pachauri, A.
Reisinger (Eds.), , 2007, p. 104, (Geneva, Switzerland).
[180] B.E. Casareto, M.P. Niraula, H. Fujimura, Y. Suzuki, Effects of carbon dioxide on
the coccolithophorid Pleurochrysis carterae in incubation experiments, Aquatic
Biology 7 (2009) 5970.
[181] D.A. Hutchins, Oceanography: forecasting the rain ratio, Nature 476 (2011)
4142.
[182] N. Aresta, I. Tommasi, M. Galatola, Potential of coring with biomass in Italy,
Energy Conversion and Management 38 (1997) S557S562.
[183] R.D. Fox, Spirulina, real aid to development, Hydrobiologia 15 (152) (1987)
9597.
[184] G.C. Clement, C. Giddey, R. Menzi, Amino acid composition and nutritive value of
alga Spirulina maxima, Journal of the Science of Food and Agriculture 18 (1967)
497501.
[185] L.H. Princen, J.A. Rothfus, Development of new crops for raw materials for industry,
Journal of the American Oil Chemists' Society 61 (1984) 281289.
[186] J.B.M. Rattray, Biotechnology and the fats and oil industry, Journal of the American
Oil Chemists' Society 61 (1984) 17011712.
[187] W. Yongmanitchai, O.P. Ward, Growth of omega-3 fatty acid production by
Pheadactylum tricornutum under different culture conditions, Applied and
Environmental Microbiology 57 (1991) 125419.
[188] E. Grima, A.R. Medina, G. Gimenez, Recovery of algal PUFAs, In: Z. Cohen (Ed.),
Chemical from Microalgae, Taylor and Francis, Philadelphia, 1999, pp. 108144.
[189] L. Hannson, M. Dostalek, Lipid formation by Cryptococcus albidus in nitrogenlimited and carbon-limited chemostat cultures, Applied Microbiology and Biotechnology 24 (1986) 187192.
[190] G. Turcotte, N. Kosaric, Lipid biosynthesis in oleaginous yeasts, Advances in
Biochemical Engineering/Biotechnology 40 (1989) 7392.
[191] S.S. Radwan, Sources of C20-polyunsaturated fatty acids for biotechnological use,
Applied Microbiology and Biotechnology 35 (1991) 421430.
[192] J.C. Nevenzel, Biogenic hydrocarbons of marine organisms, In: R.G. Ackmann
(Ed.), Marine Biogenic Lipids, Fats and Oils, vol. I, CRC Press, Florida, 1989,
pp. 372.
[193] H.R. Knapp, Omega-3 fatty acids, endogenous postaglandins and blood pressure
regulation in humans, Nutrition Review 47 (1989) 301313.
[194] D. Kyle, P. Behrens, S. Bingham, K. Arnett, D. Leiberman, Microalgae as a source
of EPA-containing oils, In: T.H. Applewhite (Ed.), World Conference on
Biotechnology for Fats and Oil Chemistry, American Oil Chemists Society, Champaign,
1990, pp. 117122.
[195] E. Fernandez, W.M. Balch, E. Maranon, P.M. Holligan, High rates of lipid biosynthesis in cultured, mesocosm and coastal populations of the coccolithophorids
Emiliania huxleyi, Marine Ecology Progress Series 114 (1994) 1322.
[196] C.P. Liu, L.P. Lin, Ultrastructural study and lipid formation of Isochrysis sp.
CCMP1324, Botanical Bulletin of Academia Sinica 42 (2001) 207214.
[197] W.L. Murphy, P.B. Messersmith, Compartmental control of mineral formation:
adaptation of a biomineralization strategy for biomedical use, Polyhedron 19
(2000) 357363.
[198] U. Riebesell, A.T. Revill, D.G. Holdsworth, J.K. Volkman, The effects of varying CO2
concentration on lipid composition and carbon isotope fractionation in Emiliania
huxleyi, Geochemisty and Cosmology Acta 64 (2000) 41794192.
[199] J.F. Rontani, D. Marchand, J.K. Volkman, NaBH4 reduction of alkenones to the
corresponding alkenols: a useful tool for their characterisation in natural
samples, Organ Geochemistry 32 (2001) 13291341.
[200] M.H. Conte, J.K. Volkman, G. Eglinton, Lipid biomarkers of the haptophyta, In: J.C.
Green, B.S.C. Leadbeater (Eds.), The Haptophyte Algae, Clarendon Press, Oxford,
1994, pp. 351377.
[201] P. Ghosh, G.W. Paterson, G.H. Wilfors, Sterols of some marine Prymnesiophyceae,
Journal of Phycology 34 (1996) 511514.
[202] I.T. Marlowe, J.C. Green, A.C. Neal, S.C. Brassell, G. Eglinton, P.A. Course, Long
chain (n-C37C39) alkenones in the Prymnesiophyceae. Distribution of
alkenones and other lipids and their taxonomic signicance, British Phycological
Journal 19 (1984) 203216.

N.R. Moheimani et al. / Algal Research 1 (2012) 120133


[203] H. Takenaka, Y. Yamaguchi, S. Teramato, N. Tanaka, M. Hori, H. Seki, et al.,
Evaluation of the mutagenic properties of the coccolithophore Pleurochrysis
carterae (Haptophyceae) as a potential human food supplement, Journal of
Applied Phycology 8 (1996) 13.
[204] H. Takenaka, Y. Yamaguchi, S. Teramato, N. Tanaka, M. Hori, H. Seki, et al., Safety
evaluation of Pleurochrysis carterae as a potential food supplement, Journal of
Marine Biotechnology 3 (1996) 274277.
[205] E. Miyamoto, F. Watanabe, S. Ebara, S. Takenaka, H. Takenaka, Y. Yamaguchi,
et al., Characterization of a vitamin B-12 compound from unicellular
coccolithophorid alga (Pleurochrysis carterae), Journal of Agricultural and Food
Chemistry 49 (2001) 34863489.
[206] S.I. Stupp, P.V. Braun, Molecular manipulation of microstructures: biomaterials,
ceramics, and semiconductors, Science 277 (1997) 12421248.
[207] A. Rames, D.F. Williams, Relationship between chemotaxis and complement
activation by ceramic biomaterials, Biomaterials 12 (1991) 661667.

133

[208] A. Piatelli, G. Podda, A. Scarano, Clinical and histological results in alveolar ridge
enlargement using coralline calcium carbonate, Biomaterials 18 (1997)
(623-7w).
[209] R. Laguna, J. Romo, B.A. Read, T.M. Wahlund, Induction of phase variation events
in the life cycle of the marine coccolithophorid Emiliania huxleyi, Applied and
Environment Microbiology 67 (2001) 38243831.
[210] D. Walsh, S. Mann, Fabrication of hollow porous shells of calcium carbonate
from self-organizing media, Nature 377 (1995) 320322.
[211] P. Van Der Wal, L. De Long, P. Westbroek, W.C. De Bruijn, Calcication in the
Coccolithophorid Alga Hymenomonas-Carterae, In: R. Hallberg (Ed.), 1983,
pp. P251P258.
[212] D.E. Outka, D.C. Williams, Sequential coccolith morphogensis in Hymnemonas
carterae, The Journal of Protozoology 18 (1971) 285297.

Das könnte Ihnen auch gefallen