Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Effect of CO2-induced seawater acidification on growth, photosynthesis and inorganic carbon acquisition of the harmful bloom-forming marine microalga, Karenia mikimotoi

  • Shunxin Hu ,

    Contributed equally to this work with: Shunxin Hu, Bin Zhou

    Roles Conceptualization, Writing – original draft

    Affiliation Department of Marine Ecology, College of Marine Life Sciences, Ocean University of China, Qingdao, China

  • Bin Zhou ,

    Contributed equally to this work with: Shunxin Hu, Bin Zhou

    Roles Writing – original draft, Writing – review & editing

    Affiliation Department of Marine Ecology, College of Marine Life Sciences, Ocean University of China, Qingdao, China

  • You Wang,

    Roles Writing – review & editing

    Affiliation Department of Marine Ecology, College of Marine Life Sciences, Ocean University of China, Qingdao, China

  • Ying Wang,

    Roles Methodology, Writing – original draft

    Affiliation Department of Marine Ecology, College of Marine Life Sciences, Ocean University of China, Qingdao, China

  • Xinxin Zhang,

    Roles Formal analysis

    Affiliation Department of Marine Ecology, College of Marine Life Sciences, Ocean University of China, Qingdao, China

  • Yan Zhao,

    Roles Writing – review & editing

    Affiliation Department of Marine Ecology, College of Marine Life Sciences, Ocean University of China, Qingdao, China

  • Xinyu Zhao,

    Roles Writing – original draft

    Affiliation Department of Marine Ecology, College of Marine Life Sciences, Ocean University of China, Qingdao, China

  • Xuexi Tang

    Roles Funding acquisition, Project administration, Writing – review & editing

    tangxx@ouc.edu.cn

    Affiliation Department of Marine Ecology, College of Marine Life Sciences, Ocean University of China, Qingdao, China

Abstract

Karenia mikimotoi is a widespread, toxic and non-calcifying dinoflagellate, which can release and produce ichthyotoxins and hemolytic toxins affecting the food web within the area of its bloom. Shifts in the physiological characteristics of K. mikimotoi due to CO2-induced seawater acidification could alter the occurrence, severity and impacts of harmful algal blooms (HABs). Here, we investigated the effects of elevated pCO2 on the physiology of K. mikimotoi. Using semi-continuous cultures under controlled laboratory conditions, growth, photosynthesis and inorganic carbon acquisition were determined over 4–6 week incubations at ambient (390ppmv) and elevated pCO2 levels (1000 ppmv and 2000 ppmv). pH-drift and inhibitor-experiments suggested that K. mikimotoi was capable of acquiring HCO3-, and that the utilization of HCO3- was predominantly mediated by anion-exchange proteins, but that HCO3- dehydration catalyzed by external carbonic anhydrase (CAext) only played a minor role in K. mikimotoi. Even though down-regulated CO2 concentrating mechanisms (CCMs) and enhanced gross photosynthetic O2 evolution were observed under 1000 ppmv CO2 conditions, the saved energy did not stimulate growth of K. mikimotoi under 1000 ppmv CO2, probably due to the increased dark respiration. However, significantly higher growth and photosynthesis [in terms of photosynthetic oxygen evolution, effective quantum Yield (Yield), photosynthetic efficiency (α), light saturation point (Ek) and ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco) activity] were observed under 2000 ppmv CO2 conditions. Furthermore, elevated pCO2 increased the photo-inhibition rate of photosystem II (β) and non-photochemical quenching (NPQ) at high light. We suggest that the energy saved through the down-regulation of CCMs might lead to the additional light stress and photo-damage. Therefore, the response of this species to elevated CO2 conditions will be determined by more than regulation and efficiency of CCMs.

Introduction

Ocean acidification refers to the ongoing reduction in the ocean pH over an extended period of time, which is primarily caused by the uptake of anthropogenic CO2 from the atmosphere [1, 2]. Industrialization and fossil fuel combustion have increased the atmospheric CO2 concentrations from pre-industrial levels of approximately 280 ppmv to the current level of approximately 390 ppmv [2, 3]. The atmospheric CO2 concentrations are predicted to increase to 1000 ppmv and 2000 ppmv by the years 2100 and 2300, respectively, if the present energy utilization structure persists [1]. Such increases in CO2 would lead to a reduction in pH (0.4 and 0.77 units, respectively) and cause substantial chemical changes in seawater carbonate systems, including increases in pCO2, HCO3- and DIC and decreases in H+ and CO32- [4, 5].

Marine phytoplankton assimilates inorganic carbon and fixes CO2 into carbohydrates through the enzyme ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco), which can only use CO2 as substrate for the carboxylase reaction. Rubisco is generally characterized by low affinities for CO2 (KM of 20–70 μmol L-1), and a competitive reaction with O2 further reduces its efficiency [6]. Therefore, photosynthesis of some marine phytoplankton might suffer from CO2 limitation, due to the present concentration of aqueous CO2 in seawater ranging from 8 to 20 μmol L-1. Most marine phytoplankton have developed so-called CO2 concentrating mechanisms (CCMs) to overcome the deficiencies of Rubisco. Two main types of CCMs are suggested to be involved in marine phytoplankton. Firstly, the dehydration of HCO3- is catalyzed by external carbonic anhydrase, facilitating the supply of CO2 at plasma membrane and improving the potential for CO2 uptake, and then CO2 accumulation is achieved by the active transport of HCO3- or CO2 at the chloroplast envelope. Secondly, HCO3- is transported across the plasmalemma and/or chloroplast envelope and then converted to CO2, catalyzed by one of several types of internal carbonic anhydrase [7]. However, the expression and operation of CCMs require energetic investment. Therefore, marine phytoplankton could benefit from elevated pCO2 in two synergetic ways: one is that the increased dissolved carbon dioxide (CO2aq) would supply additional substrate for photosynthetic carbon fixation and reduce the oxygenase reaction of Rubisco, alleviating the carbon limitations of species without a CCM [8]. The other is that, increased pCO2 could down-regulate the energetically costly operation of CCMs.

Many marine phytoplankton species down-regulate their operation of CCMs at high pCO2 conditions, as revealed by a lower photosynthetic affinity for CO2, decreased activities of carbonic anhydrase and/or a lower contribution of HCO3-assimilation [912]. This is taken as evidence that elevated pCO2 exerts positive effects on the growth and photosynthesis of species bearing CCMs[9, 1315]. Some species (Macrocystis pyrifera and Gracilaria lemaneiformis), however, do not show any deactivation of CCMs or changes in growth and photosynthesis in response to elevated pCO2 [16, 17]. In addition, increased pCO2 might exert negative effects on calcifying species because of a lowering of saturation of CaCO3, which might make calcfication more difficult [18,19]. Deleterious effects of elevated pCO2, however, also occur on non-calcifying species [2023], probably due to the negative effects on physiological processes caused by reduced external pH. [24]. Altogether, these findings havedeepened our understandings of differential physiological responses to elevated pCO2 among various marine phytoplankton species, such as diatoms and coccolithophores.

Karenia mikimotoi is a widespread, toxic and non-calcifying dinoflagellate, which can have deleterious effects on other marine phytoplankton, fish and shellfish through the release of ichthyotoxins and hemolytic toxins [25]. K. mikimotoi toxic blooms have been reported in Japan, Ireland, England, France, India and China [2630]. At least 103 harmful algal blooms (HABs) have been reported between 2004 and 2014 caused by K. mikimotoi in eastern and southern Chinese seas, affecting about 37527 km2 (according to China Marine Disaster Bulletin). Owing to its ecological and environmental implications, it is necessary to study the physiological responses of this species to elevated pCO2 in order to predict the occurrence, severity and impacts of blooms of this species in the future. In the present study, we evaluated growth, photosynthesis, dark respiration and the CCMs modes of K. mikimotoi exposed to three different pCO2 levels: 390 ppmv (pHNBS: 8.10) which is the present pH value, as well as 1000 ppmv (pHNBS: 7.78) and pCO2: 2000 ppmv (pHNBS: 7.49), which are predicted to be possible conditions in 2100 and 2300, respectively. We hypothesized that (1) the operation of CCMs will be down-regulated with elevated pCO2 and (2) the reduction in the energy costs of CCMs will benefit growth and photosynthesis of K. mikimotoi.

Materials and methods

Culture conditions and experimental design

Karenia mikimotoi (strain OUC151001) was obtained from the Algal Culture Collection at the Ocean University of China. Cells were cultured in 0.45μm-filtered natural seawater, collected from Luxun Seaside Park (Qingdao), which had been autoclaved (30min, 121°C) and enriched with f/2 medium [31]. All cultures were incubated at 20±1°C and illuminated with 80 μmol photon m-2 s-1 under a 12:12 light: dark cycle.

Experiments were conducted in triplicate 1000 ml sterilized and acid-washed Erlenmeyer flask containing 600 ml of medium. Prior to inoculation, the cultures were equilibrated at three different CO2 levels: 390 ppmv CO2 (~present-day), 1000 and 2000 ppmv CO2 (predicted CO2 levels in 2100 and 2300, respectively), obtained by gentle bubbling with 0.22 μm-filtered ambient air and air/CO2 mixtures. The air/CO2 mixtures were generated by plant CO2 chambers (HP400G-D, Ruihua Instrument & Equipment Ltd, Wuhan, China) with a variation of less than 5%. Semi-continuous cultures were used to measure the effects of CO2-induced seawater acidification on the growth and physiology of K. mikimotoi in the present study, similar to previous ocean acidification research [9, 3234]. All cultures were diluted to 800 cells mL-1 with fresh medium pre-acclimated to the desired CO2 level every 24h to maintain cells in exponential growth phase, and to minimize pH fluctuations. Cultures were harvested following 4–6 weeks of semi-continuous incubation when the growth rates were not significantly different for three or more consecutive days, which was considered fully acclimated to their respective experimental treatments.

Seawater carbonate chemistry

The pH value and dissolved inorganic carbon (DIC) were determined prior to and after the daily dilution. The concentration of DIC in the culture medium was measured using a total organic carbon analyzer (TOC-VCPN, Shimadzu). The samples were filtered onto brown glasses via 0.45 μm cellulose acetate membranes and stored in a refrigerator (4°C). The pH was measured using a pH meter (SevenCompact S210k, Mettler Toledo, Switzerland), which was calibrated daily with standard National Bureau of Standards (NBS) buffer system. The other relevant parameters of carbonate system were determined with the CO2SYS software [35], based on the known parameters (pH, DIC, salinity and temperature).

Growth and elemental analysis

Cell growth rate was monitored daily using a plankton counting chamber (0.1 mL) before and after the medium was diluted. The specific growth rate (μ) was calculated using the equation:

where N0 and N1 represent the average cell numbers at times t0 (after the dilution) and t1 (before the dilution), respectively.

Samples for measurements of cellular carbon (C), nitrogen (N) and phosphorus (P) were filtered onto glass microfiber membranes (GF/F, Whatman), which were pre-combusted at 500°C for 4 h, and then stored at -20°C in a refrigerator before analysis. Cellular C and N content were determined using a Perkin-Elmer 2400 CHNS analyzer following the method of Zhao et al. [36]. Cellular P content was measured as in Fourqurean et al. [37].

Chlorophyll a

Chlorophyll a (Chl a) samples from the cultures were filtered onto glass microfiber filters (GF/F, Whatman), and extracted with 10 mL of methanol overnight at 4°C. Samples were then analyzed in a spectrophotomer, and the Chl a concentration was calculated according to the following equation [38]: where A652, A665 and A750 denoted the absorbance values of the methanol extracts at 652nm, 665nm and 750nm, respectively.

Photosynthetic oxygen evolution and respiration

Photosynthetic oxygen evolution was measured under the same light intensities used for growing cultures, with lighting provided by a halogen lamp. The dark respiration rate was determined using a Clark-type oxygen electrode (Chlorolab 3, Hansatech, UK). Experimental temperature was maintained at 20°C using a water bath circulator. Before the determination, cells were acclimated to light or dark conditions in the reaction chamber for 20 min. The 5ml-reaction media was continuously stirred with a magnetic stirrer during treatment. DIC concentrations were consistent with their culture conditions, which were nominally 390ppmv: 1919–1924 μmol/kg; 1000ppmv: 2059–2067 μmol/kg; 2000 ppmv:2152–2156 μmol/kg, respectively.

Chlorophyll fluorescence measurements

Fluorescence induction curves and rapid light curves (RLCs) were applied to evaluate the photosynthetic performance of K. mikimotoi acclimated to different pCO2, using a Water-PAM fluorometer.

The RLCs were measured at 8 different actinic irradiance levels (80, 119, 184, 276, 393, 546, 897 and 1315 μmol photon m-2 s-1), each of which lasted 10s. To quantitatively compare the RLCs of K. mikimotoi acclimated to different pCO2, the equation of Platt et al. [39] was applied to derive characteristic parameters: photosynthetic efficiency (α), light saturation point (Ek), photo-inhibition rate of photosystemII(β) and maximum relative electron transport rate (rETRmax). The light saturation point was determined from: Ek = rETRmax/α [40]. Related parameters were applied to determine the convergence of the regression mode according to Ralph and Gademann [40].

For the fluorescence induction curves, all of the samples were dark-acclimated for 20 mins before determination. The dark-acclimation induction curves were measured with a delay of 40 s between the determinations of Fv / Fm. The actinic light was set at 80, 276 and 897 μmol photon m-2 s-1, respectively, to measure the value of effective quantum yield (Yield) and non-photochemical quenching (NPQ)

Determination of ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco) activities

Cells were collected by centrifugation at a rotating speed of 3000g at 4°C for 15 min. After removing the supernatant, cells were grinded on ice with the addition of 1mL buffer solution (40 mM Tris-HCl, 5 mM Glutathione, 10 mM MgCl2 and 0.25 mM EDTA, pH 7.6). The liquid was concentrated, and the supernatant was used for further assays. Rubisco activity in the supernatant was generally determined following the methods described by Gerard and Driscoll [41]. The assay mixture contained 5 mM NADH, 50 mM ATP, 50 mM phosphocreatine, 0.2 mM NaHCO3, 160 U/mL creatine phosphokinase, 160 U/mL Phosphoglycerate kinase, 160 U/mL glyceraldehyde-3-phosphate dehydrogenase and reaction buffer (0.1M Tris-HCl, 12 mM MgCl2 and 0.4 mM EDTA, pH 7.8). Absorbance values at 340 nm (A340) were measured every 20s for 3 min to obtain the background NADH oxidation rate. A volume of 0.05 mL of RuBP (final concentrations of 25 mM) was added into the assay mixture, and the A340 was then recorded every 20 s for 3 min. The activities of Rubisco were computed by subtracting the background rate of decrease in A340 from the rate determined during the three minutes following RuBP addition, and then converting the corrected rate of A340 decrease to a rate of NADH oxidation.

pH drift experiment

A pH drift experiment was applied to determine whether K. mikimotoi can utilize HCO3- as an inorganic carbon source; the ability of algae to raise the medium pH to higher than 9.0 is considered as evidence of its capacity to utilize HCO3-. The experiment was performed in sterilized glasses containing 10 mL samples (cell concentrations of 10×104 mL-1) at 20°C and 80 μmol photon m-2 s-1. The pH was measured every hour and the final pH values were obtained once no further pH increases were detected.

Determination of carbonic anhydrase activity

Cells were collected by centrifugation at a rotating speed of 3000g at 4°C for 15 min and re-suspended in 20mM barbitone (pH 8.2). The total carbonic anhydrase (CAtot) and external carbonic anhydrase (CAext) activities were measured using an electrometric method [42]. For determination of CAtot, cells were disrupted with a sonicator, and cell brakage confirmed under a microscope. The reaction was begun by adding 2 mL ice-cold CO2 to saturated Milli-Q water, and the time it took for the pH to decrease from 8.2 to 7.2 was recorded. The temperature was controlled at 4°C.

Effect of inhibitors on chlorophyll fluorescence

The RLCs of K. mikimotoi acclimated to different pCO2 with addition of inhibitors were obtained to determine the mechanism of inorganic carbon acquisition. The inhibitors included acetazolamide (AZ), which inhibits only extracellular CA, ethoxyzolamide (EZ), which inhibits both extracellular and intracellular CA, and DIDS (4,4’-diisothiocyanostilbene-2,2’-disulfonate), which inhibits direct HCO3- uptake by means of the anion-exchange protein. These inhibitors have been widely used to determine the contribution of external CA, internal CA and anion-exchange protein to photosynthetic inorganic carbon uptake [21, 43, 44].

Statistical analysis

One-way ANOVA was used to analyze the significance of the differences between treatments using the SPSS software (20.0), and the significant difference level was set to P < 0.05. All figures were prepared with Sigmaplot 12.5.

Results

Seawater carbonate chemistry

Under the simulated laboratory conditions of ocean acidification, the seawater carbonates chemistry system at elevated pCO2 (1000 ppmv and 2000 ppmv) levels significantly differed from that of the control group (Table 1). The DIC, CO2 and HCO3- concentrations in the 1000 ppmv-treated system were increased by 7.0%, 125.9% and 10.1%, respectively, whereas in the 2000 ppmv-treated system, these values increased by 12.1%, 361.4% and 15.5%, respectively. The CO32- concentrations were decreased by 46.5% and 71.1% in the 1000 ppmv- and 2000 ppmv-treated systems, respectively, and the difference in total alkalinity (TA) was insignificant. The fluctuations of pH prior and after the dilution of the culture medium were <0.03.

thumbnail
Table 1. Parameters of the seawater carbonate chemistry system at different pCO2 levels prior and after the dilution. The dissolved inorganic carbon (DIC) concentration, pHNBS, temperature and salinity were used to compute other parameters with a CO2 system analyzing software (CO2SYS). Data are shown as the mean ± SE (n = 9). Different letters represent significant difference between variables (P < 0.05).

https://doi.org/10.1371/journal.pone.0183289.t001

Growth and elemental composition

The growth rates at the three pCO2 levels are shown in Fig 1. The growth rate of K. mikimotoi was significantly stimulated by 16.84% (P<0.05) after exposure to 2000 ppmv pCO2; although growth was also enhanced under 1000 ppmv pCO2, the increase was not statistically significant (P>0.05).

thumbnail
Fig 1. Growth rate of K. mikimotoi acclimated to different pCO2 levels. Data are shown as the mean ± SE (n = 9).

https://doi.org/10.1371/journal.pone.0183289.g001

The total cellular C,N,P and their elemental ratio in K. mikimotoi are shown in Table 2. The cellular C and P concentrations of K. mikimotoi exposed to 2000 ppmv pCO2 levels were significantly (P<0.05) higher than those of the control, whereas there was no significant difference between the control and 1000 ppmv pCO2 (P>0.05). Furthermore, elevated pCO2 exerted no significant effects on the cellular N, C: N, C: P and N: P ratios (P>0.05).

thumbnail
Table 2. Elemental composition (total C, N and P) and elemental ratio (C:N, C:P and N:P) of K. mikimotoi acclimated to different pCO2 levels. Data are shown as the mean ± SE (n = 9).

Different letters represent significant difference between variables (P < 0.05).

https://doi.org/10.1371/journal.pone.0183289.t002

Chlorophyll a

Cells acclimated to the three pCO2 levels showed the same Chl a content (Fig 2) of about 2.0 pg cell-1, with no significant differences among treatments.

thumbnail
Fig 2. Chlorophyll a content of K. mikimotoi acclimated to different pCO2 levels. Data are shown as the mean ± SE (n = 9).

https://doi.org/10.1371/journal.pone.0183289.g002

Photosynthetic oxygen evolution, dark respiration and Rubisco activities

Results of the determinations of the net photosynthetic oxygen evolution, gross photosynthetic oxygen evolution, dark respiration, and Rubisco activity are shown in Fig 3. Net photosynthetic oxygen evolution (Fig 3A) and Rubisco activity (Fig 3D) were significantly enhanced by 22.86% (P<0.05) and 31.99% (P<0.05) under 2000 ppmv pCO2, whereas there were no significant difference between the control and 1000 ppmv pCO2 (P>0.05). Cells acclimated to both 1000 ppmv and 2000 ppmv pCO2 treatments showed higher gross photosynthetic oxygen evolution (Fig 3B) and dark respiration (Fig 3C) than those of the control (P<0.05).

thumbnail
Fig 3. Net photosynthetic oxygen evolution (A), gross photosynthetic oxygen evolution (B), dark respiration (C) and Rubisco activity (D) of K. mikimotoi acclimated to different pCO2 levels. Data are shown as the mean ± SE (n = 9).

https://doi.org/10.1371/journal.pone.0183289.g003

Chlorophyll fluorescence

Rapid light curves (RLCs) were determined at the three levels of pCO2 (control, 1000 ppmv and 2000 ppmv). The three treatments all exhibited a classical pattern of rETR as a function of PAR (Fig 4A), with a rapid increase under light-limited conditions followed by a plateau, at which the photosynthetic pathway was saturated.

thumbnail
Fig 4. The rapid light curves of K. mikimotoi without and with the addition of inhibitors (AZ, EZ and DIDS) acclimated to different pCO2. Data are shown as the mean ± SE (n = 3).

https://doi.org/10.1371/journal.pone.0183289.g004

The parameters derived from the RLCs are shown in Table 3. The photosynthetic efficiency (α) of K. mikimotoi was significantly stimulated by 23.1% (P<0.05) under 2000 ppmv pCO2. Cells acclimated to 1000 ppmv pCO2 also showed higher α than that of control, but the increase was not statistically significant (P>0.05). Contrary to the trend observed in α, elevated pCO2 significantly decreased the light saturation point (Ek) of K. mikimotoi by 7.5% (P<0.05) and 10.2% (P<0.05) under 1000 and 2000 ppmv CO2 compared with the control, respectively. The maximum relative electron transport rates (rETRmax) were approximately 10% higher in both elevated pCO2 levels compared with the control, but the differences were not significant (P > 0.05). Moreover, the results showed that increased pCO2 levels had no significant effect on the Fv / Fm of K. mikimotoi (P > 0.05).

thumbnail
Table 3. Photosynthetic parameters derived from the rapid light curves of K. mikimotoi acclimated to different pCO2 levels. Data are shown as the mean ± SE (n = 3).

Different letters represent significant difference between variables (P < 0.05).

https://doi.org/10.1371/journal.pone.0183289.t003

The inhibition of rETR was obtained from the rapid light curves in all three different pCO2 levels when exposed to the actinic irradiance of 1315 μmol photon m-2 s-1. β values (relative inhibition), characterizing the photo-inhibition rate of PSII exposed to high actinic irradiance, were significantly increased by 126.7% (P<0.01) and 194.3% (P<0.001), respectively, in 1000 ppmv and 2000 ppmv pCO2 compared with the control (Table 3).

The non-photochemical quenching (NPQ) and effective quantum yield of PS II (Yield) values derived from the induction light curve (IC) are shown in Figs 5 and 6, respectively. The results indicated that response of NPQ to higher CO2 concentrations depended on actinic irradiances. Under 80 μmol photon m-2 s-1, the NPQ values in the 390ppmv pCO2 were 15.4% (P<0.05) and 43.9% (P<0.01) higher than those observed in the 1000 ppmv and 2000 ppmv pCO2, respectively. By contrast, NPQ values at 276 and 897 μmol photon m-2 s-1 were significantly increased in the two high pCO2 groups (P < 0.01), but there was no significant difference between 1000 ppmv and the control when exposed to 276 μmol photon m-2 s-1 (P>0.05). Relative to the control conditions, both the 1000 ppmv and 2000 ppmv pCO2 groups significantly stimulated the Yield, which was increased by 4.0% (P<0.05) and 4.9% (P<0.01) at 80 μmol photon m-2 s-1, and increased by 11.8% (P<0.05) and 11.8% (P<0.05) at 897 μmol photon m-2 s-1.

thumbnail
Fig 5. Non-photochemical quenching (NPQ) of K. mikimotoi acclimated to different pCO2 levels at an actinic irradiance of 80, 276 and 897 photon m-2 s-1. Data are shown as the mean ± SE (n = 3).

https://doi.org/10.1371/journal.pone.0183289.g005

thumbnail
Fig 6. Effective quantum yield (Yield) of K. mikimotoi acclimated to different pCO2 levels at an actinic irradiance 80, 276 and 897 photon m-2 s-1. Data are shown as the mean ± SE (n = 3).

https://doi.org/10.1371/journal.pone.0183289.g006

At an actinic of 276 μmol photon m-2 s-1, the 2000 ppmv pCO2 groups exhibited a significant (P<0.01) increase, but not the 1000 ppmv pCO2 groups.

pH drift experiment and carbonic anhydrase activity

The final pH value obtained in the pH drift experiment was 9.8±0.1. Furthermore, the total carbonic anhydrase activity (CAtot) and internal anhydrase activity (CAint) (Fig 7) of K. mikimotoi were significantly decreased by 50.6% (P<0.01) and 55.5% (P<0.05) after exposure to 2000 ppmv pCO2, and no significant changes were observed when exposed to 1000 ppmv pCO2 (P>0.05). The external carbonic anhydrase activity (CAext) (Fig 7) of K. mikimotoi was significantly lower than the CAint, and no significant changes were observed when exposed to 1000 ppmv (P>0.05) and 2000 ppmv (P>0.05) pCO2.

thumbnail
Fig 7. Total, external and internal carbonic anhydrase activity of K. mikimotoi acclimated to different pCO2 levels. Data are shown as the mean ± SE (n = 3).

https://doi.org/10.1371/journal.pone.0183289.g007

Effect of inhibitors on chlorophyll fluorescence

The rapid light curves of K. mikimotoi acclimated to different pCO2 levels with and without the addition of AZ, EZ and DIDS are shown in Fig 4. The results indicated that the rETR values of K. mikimotoi were significantly inhibited by the addition of EZ and DIDS, whereas the inhibition of rETR by the addition of AZ was significantly less than that obtained with EZ and DIDS.

The inhibition rates of rETR at different pCO2 levels obtained with the addition of AZ, EZ and DIDS are shown in Table 4. The results indicated that the inhibitions of rETRs by the addition of EZ and DIDS were significantly (P<0.01) higher in the controls than those obtained at the high pCO2 groups (1000 ppmv and 2000 ppmv). Furthermore, the results from all three pCO2 levels indicated that the inhibition of rETR was significantly increased by the increasing PAR. in the presence of AZ, rETR inhibition was only observed in the 390 ppmv pCO2.

thumbnail
Table 4. Percent inhibition of rETR acclimated to different pCO2 with the addition of AZ, EZ and DIDS within a PAR range of 0 to 1315 μmol photon m-2 s-1, “—” represents no inhibition of rETR. Data are shown as the mean ± SE (n = 3).

https://doi.org/10.1371/journal.pone.0183289.t004

Discussion

Inorganic carbon acquisition

Dinoflagellates are abundant and ecologically important in marine ecosystems, and morphologically and physiologically diverse [45]. They are also the only oxygenic photoautotrophs with type II Rubisco, the enzyme with the lowest affinity for CO2 among eukaryotic phytoplankton. Thus, dinoflagellates are at a disadvantage with regard to photosynthetic carbon fixation under the present ocean conditions of low CO2, and high O2 [6]. Consequently, dinoflagellates probably require an efficient CCM to compete with other phytoplankton that have higher photosynthetic and growth rates. It is a common notion that the ability of algae to raise the final pH of the medium to higher than 9.0 is an indicator of HCO3- utilized by the species [46, 47]. The pH-drift experiments conducted in our study indicated that the final pH value in the medium of K. mikimotoi was 9.8±0.1, suggesting that K. mikimotoi could use HCO3- in seawater. Although most phytoplankton species possess CCMs, large differences exist in their efficiencies. Owing to their rather inefficient CCMs (strongly dependent on CO2 as inorganic carbon source), the photosynthetic carbon fixation of the coccolithophorid Emiliania huxleyi and the raphidophyceae Heterosigma akashiwo are well below saturation at present CO2 levels, and therefore are more CO2-sensitive than species with highly efficient CCMs (which rely heavily on HCO3- as inorganic carbon source), such as Skeletonema costatum and Phaeocystis globose [32, 48]. In the present study, Rubisco activity of K. mikimotoi did not change between 390 and 1000 ppmv, and ETR values only increase slightly (~10%), which is strong evidence that K. mikimotoi, similar to S. costatum and P. globose, relies heavily on HCO3- as an inorganic carbon source.

CAext, which catalyzes the dehydration of HCO3- to CO2 at the cell surface [4749], was found to decrease at high pCO2 conditions in other species such as Phaeocystis globosa and S. costatum, underlining the important role of CAext in inorganic carbon acquisition [48]. In the present study, such a HCO3- dehydration mechanism was likely to contribute very little to the inorganic carbon acquisition in K. mikimotoi, because the activities of CAext were very low in all treatments (Fig 7). Furthermore, the minor role of CAext was also indicated by the addition of membrane impermeable CA inhibitor AZ, which only slightly inhibited the rETR of K. mikimotoi (Fig 4 and Table 4). The low CAext activity observed in K. mikimotoi was consistent with most other tested dinoflagellate species. Rost et al. [50] investigated CAext activities in Prorocentrum minimum, Heterocapsa triquetra and Ceratium lineatum, and found relatively low or negligible activities in all three species. Eberlein et al. [51] showed that CAext activities of the dinoflagellate Alexandrium tamarense acclimated to a range of pCO2 from 180 to 1200 μatm were close to detection limits, and thus only played a minor role. However, low CAext activity is not universal in all dinoflagellate species. High activity of CAext was found in Scrippsiella trochoidea, probably to convert the effluxing CO2 to HCO3-, and then utilized via the HCO3- transporter by the cells [51]. This ‘CO2 recirculation mechanism’ might be especially beneficial for species with high dark respiration rates.

Furthermore, direct HCO3- uptake via the anion-exchange (AE) protein has also been observed in other species, which suggests that HCO3- utilization could be inhibited by the AE protein inhibitor, DIDS [5253]. From our results, such a direct HCO3- uptake was likely to be present in K. mikimotoi, and it was significantly reduced when the cells acclimated to 1000 and 2000 ppmv pCO2, because the rETR of K. mikimotoi was drastically depressed by the addition of DIDS, and the inhibition decreased at high pCO2 as compared with the control (Fig 4 and Table 4). Down-regulated CCMs at high pCO2 have also been found in widely distributed species such as Skeletonema costatum, Emiliania huxleyi, Thalassionema nitzschioides and Pseudo-nitzschia multiseries [48,54,55]. This down-regulation might result from the increasing diffusive CO2 uptake at high pCO2 conditions, since CO2 uptake is considered to be less energetically costly than HCO3- uptake. Consequently, cells can optimize their allocation of energy and apportion more energy for photosynthetic carbon fixation. Marine phytoplankton productivity based on energy or carbon content might thus increase under typically resource-limited conditions in the ocean. From this point of view, species with regulated CCMs, as shown for K. mikimotoi, might have a competitive advantage in the future compared to species that do not react to high pCO2 such as Phaeocystis globosa, Thalassiosira pseudonana, Eucampia zodiacus and Nitzschia navis-varingica [48,54,55]. As a consequence, these different responses of CCMs to elevated pCO2 might change the fitness of the different group and possibly alter the distribution and succession of marine algae in natural systems.

Growth, photosynthesis, respiration and photosynthetic electron transport

Growth is a comprehensive parameter integrating all physiological processes in marine phytoplankton, and different responses of growth to increased pCO2 have been reported among different marine phytoplankton species, with positive, negative and no significant responses. According to our experimental results, the growth rate of K. mikimotoi was not significantly affected under 1000 ppmv pCO2 conditions, while the stimulated growth rate was observed under 2000 ppmv pCO2 conditions. Enhanced growth and photosynthesis by elevated pCO2 have been reported in species such as the diatoms Navicula pelliculosa [56] and Phaeodactylum tricornutum [9], the raphidophyceae Heterosigma akashiwo [33] and the chlorophyte Ulva rigida [57]. However, other studies found no significant effects [32, 58, 59], or even negative effects [22, 60, 61] on the growth, photosynthesis or primary productivity of marine phytoplankton. With regard to the species with highly efficient and strongly regulated CCMs, stimulated growth by elevated pCO2 is generally attributed to decreased energetic cost of CCMs and HCO3- uptake, with the saved energy allocated to support growth [62]. For K. mikimotoi, the operation of CCMs were downregulated under both 1000 and 2000 ppmv pCO2 conditions, but the saved energy from the downregulated CCMs did not stimulate growth under 1000 ppmv pCO2, probably because of enhanced respiratory carbon loss. Even though gross photosynthesis of K. mikimotoi was enhanced under 1000 ppmv pCO2 (Fig 3B), net photosynthesis was not significantly affected (Fig 3A), which may be largely caused by the enhanced dark respiration. Consequently, the growth of K. mikimotoi grown in 1000 ppmv pCO2 was not significantly stimulated, similarly to what was reported for the diatom Thalassiosira pseudonana [10]. When the cells acclimated to 2000 ppmv pCO2, photosynthesis of K. mikimotoi significantly increased. The explanation might be that energy was saved from down-regulation of CCMS, and thus resulted in increased growth, Results in the present study suggest that the responses of marine phytoplankton to future CO2-driven seawater acidification are not only determined by the efficiency and regulation of CCMs, but also controlled by the balance of the positive and negative effects associated with increased pCO2 and seawater acidity.

Enhanced dark respiration rates by elevated pCO2 have been reported in other species such as the diatom Thalassiosira pseudonana, the dinoflagellate Alexandrium tamarense and the diatom Phaeodactylum tricornutum [9, 10, 52], but not in the dinoflagellate Scrippsiella trochoidea and the rhodophyte Porphyra leucosticte [51, 63]. Conversely, a decrease in dark respiration was observed in the chlorophyte Ulva rigida and the diatom T. pseudonana [64, 65]. Stimulation of dark respiration under high pCO2 condition could reflect higher energy requirement due to either enhanced biosynthesis in response to increased carbon fixation, more energy demand to counteract external pH reduction and to maintain intracellular acid-base stability [66], or pH-dependent changes in the function of respiratory enzymes and altered proton gradient across the mitochondrial membrane [67]. Alternatively, decreased dark respiration by elevated pCO2 has been ascribed to the down-regulation of CCMs in order to prevent oxidative damage from excess energy [65].

Comparing the parameters derived from rapid light curves (RLCs) and induction curves, elevated pCO2 appeared to have a positive impact on the efficiency of PSII, indicated by stimulated α, Yield and decreased Ek at high pCO2. Generally, photosynthetic efficiency (α) represents the energetic costs of photosynthesis. Accordingly, the present study indicated that future increased pCO2 reduces the costs of photosynthesis in K. mikimotoi. Fu et al. (2007) suggested that a stimulated α at high pCO2 is attributed to the decreased energetic cost of CCMs and more efficient light use [32]. This suggestion is supported by our findings because the lower contribution of HCO3- to inorganic acquisition was also observed in K. mikimotoi. The Ek (rETRmax / α) represents the optimum light of the photosynthetic apparatus to maintain a balance between photosynthetic energy capture and the capacity to process this energy [68]. For K. mikimotoi, an increase in α, and no significant changes in ETRmax at high pCO2 resulted in a decrease in the light saturation point (Ek). This indicates that elevated pCO2 can stimulate the efficiency of light harvesting and processing of PSII, and thus fewer photons are demanded to reach the Ek at high pCO2. Furthermore, the lower Ek of K. mikimotoi in future CO2-induced ocean acidification also suggests that light is less likely to be a limiting factor, thus making it more competitive in light-limited conditions. By contrast, the lower Ek also indicated that high pCO2 lowered the light threshold at which light became excessive in K. mikimotoi, and thus the cells easily became photo-inhibited at high light conditions, an inference supported by the increased NPQ (Fig 5) and β (Table 3) at high pCO2 conditions. The operation of CCMs serves as the pathway for alleviating photo-damage through the dissipation of excess light energy [69, 70]. For K. mikimotoi, the more active CCMs in the control would consume more energy and drain more hydrogen ion out of the thylakoid lumen to the stroma, which results in a lower NPQ (Fig 5). Therefore, elevated pCO2 diminished the energy-dissipation via the down-regulated CCMs, leading to increased NPQ (Fig 5) and enhancing photo-inhibition (Table 3) at high light conditions. Wu et al. (2010) also found that elevated pCO2 enhanced the photo-inhibition of rETR in Pheodactylum tricornutum when exposed to high PAR, but the NPQ decreased at high pCO2 [9]. In contrast, Thalsassiosira pseudonana acclimated to 390 and 1000 ppmv CO2 showed identical photo-inhibition and NPQ when exposed to high PAR, indicating that high light tolerance was not altered by high pCO2 [10]. Such different responses among K. mikimotoi, P. tricornutum and T. pseudonana suggest that species-specific metabolic pathways might be involved in coping with elevated pCO2 and high light stress.

Elemental composition

The elemental composition of marine phytoplankton differs intra and interspecifically. It has been hypothesized that elevated pCO2 could increase carbon assimilation, and thereby alter the elemental composition of marine phytoplankton. In this study, the similar activities of Rubisco (Fig 3D) and growth rate (Fig 1) under 390 and 1000 ppmv CO2 treatments suggest that photosynthetic carbon fixation did not differ between these conditions, which could explain why the total cellular carbon, nitrogen and phosphorus contents of K. mikimotoi were unaffected by the 1000 ppmv CO2 conditions. However, the total cellular carbon and phosphorus contents of K. mikimotoi increased under 2000 ppmv CO2, likely due to either enhancement of photosynthetic carbon fixation and elevated uptake of PO43-, or to increased activities of phosphatase. Furthermore, enhancement of protein and carbohydrate synthesis also likely contributed to the increased cellular C and P. The unchanged cellular nitrogen contents under 2000 ppmv CO2 were probably determined by the balance between enhanced nitrogen accumulation and nitrogen loss.

Ecological and environmental implications

It has been proposed that the dominance of bloom-forming species might be dependent on their ability to operate a regulated and efficient CCM [48,54]. In the current study, K. mikimotoi was found to have an efficient CCM, and the operation of CCM was down-regulated at high pCO2 (1000 ppmv and 2000 ppmv) conditions. However, the growth of K. mikimotoi in 1000 ppmv pCO2 was not stimulated by the reduced energetic costs of the CCM, probably due to additional carbon loss caused by enhanced dark respiration. Gao et al. [71] conducted an experiment to investigate the responses of natural phytoplankton assemblages in the South China Sea grown, over a range of light, to elevated pCO2. The results showed that growth rates of three diatom species (Thalassiosira pseudonana, Phaeodactylum tricornutum, and Skeletonema costatum) under 1000 ppmv pCO2 were significantly stimulated at low light levels. The different responses between the dinoflagellate K. mikimotoi and the diatoms T. pseudonana, P. tricornutum and S. costatum suggest that ongoing CO2-related changes could affect their dominance and succession in the future, possibly not favoring K. mikimotoi in inter-specific competitions.

Global changes not only involve increasing CO2 levels, but also other environmental factors such as shifts in light availability, nutrient supplies and temperature. Further studies will need to investigate whether species response to elevated pCO2 might be modulated by other interactive environmental factors. It will be critical to establish the environmental, ecological and economic consequences of K. mikimotoi blooms in a future changing ocean.

Acknowledgments

We are thankful to all of the members of the laboratory and Dr. Liang in the College of Fisheries in Ocean University of China for help. This study was supported by the Natural Science Foundation of China (41476091) and NSFC-Shangdong Joint Fund (U1406403)

References

  1. 1. Caldeira K, Wickett ME. Oceanography: anthropogenic carbon and ocean pH. Nature. 2003;425(6956): 365–365. pmid:14508477
  2. 2. Doney SC, Fabry VJ, Feely RA, Kleypas JA (2009) Ocean acidification: the other CO2 problem. Mar Sci 1. 2009.
  3. 3. Griggs DJ, Noguer M. Climate change 2001: the scientific basis. Contribution of working group I to the third assessment report of the intergovernmental panel on climate change. Weather. 2002;57(8): 267–269.
  4. 4. Gattuso JP, Allemand D, Frankignoulle M. Photosynthesis and calcification at cellular, organismal and community levels in coral reefs: a review on interactions and control by carbonate chemistry. Am Zool. 1999;39(1): 160–183.
  5. 5. Raven J, Ball L, Beardall J, Giordano M, Maberly SC. Algae lacking carbon-concentrating mechanisms. Can J Botany 2005;83(7): 879–890.
  6. 6. Badger MR, Andrews TJ, Whitney SM, Ludwig M, Yellowlees DC, Leggat W, et al. The diversity and coevolution of Rubisco, plastids, pyrenoids, and chloroplast-based CO2-concentrating mechanisms in algae. Can J Botany.1998;76(6): 1052–1071.
  7. 7. Nimer NA, Iglesias-Rodriguez MD, Merrett MJ. Bicarbonate utilization by marine phytoplankton species. J Phycol.1997; 33(4): 625–631.
  8. 8. Raven JA, Giordano M, Beardall J, Maberly SC. Algal evolution in relation to atmospheric CO2: carboxylases, carbon-concentrating mechanisms and carbon oxidation cycles. Philos T Roy Soc Lon B. 2012; 367(1588): 493–507.
  9. 9. Wu Y, Gao K, Riebesell U. CO2-induced seawater acidification affects physiological performance of the marine diatom Phaeodactylum tricornutum. Biogeosciences.2010; 7(9): 2915–2923
  10. 10. Yang G, Gao K. Physiological responses of the marine diatom Thalassiosira pseudonana to increased pCO2 and seawater acidity. Mar Environ Res. 2012; 79: 142–151 pmid:22770534
  11. 11. Rost B, Riebesell U, Burkhardt S, Sültemeyer D. Carbon acquisition of bloom-forming marine phytoplankton. Limnol Oceanogr. 2003; 48(1): 55–67.
  12. 12. Kranz SA, Dieter S, Richter KU, Rost B. Carbon acquisition by Trichodesmium: the effect of pCO2 and diurnal changes. Limnol Oceanogr. 2009;54(2): 548–559.
  13. 13. Kim JM, Lee K, Shin K, Kang J H, Lee HW, Kim M, et al. The effect of seawater CO2 concentration on growth of a natural phytoplankton assemblage in a controlled mesocosm experiment. Limnol oceanogr.2006;51(4): 1629–1636.
  14. 14. Olischläger M, Bartsch I, Gutow L, Wiencke C. Effects of ocean acidification on different life-cycle stages of the kelp Laminaria hyperborea (Phaeophyceae) Bot Mar. 2012; 55(5): 511–525.
  15. 15. Olischläger M, Bartsch I, Gutow L, Wiencke C. Effects of ocean acidification on growth and physiology of Ulva lactuca (Chlorophyta) in a rockpool-scenario. Phycol Res. 2013; 61(3): 180–190.
  16. 16. Zou D, Gao K. Effects of elevated CO2 on the red seaweed Gracilaria lemaneiformis (Gigartinales, Rhodophyta) grown at different irradiance levels. Phycologia. 2009; 48(6):510–517.
  17. 17. Fernández P A, Roleda M Y, Hurd C L. Effects of ocean acidification on the photosynthetic performance, carbonic anhydrase activity and growth of the giant kelp Macrocystis pyrifera. Photosynth Res. 2015; 124(3):293–304. pmid:25869634
  18. 18. Kurihara H, Asai T, Kato S, Ishimatsu A. Effects of elevated pCO2 on early development in the mussel Mytilus galloprovincialis. Aquat. Biol. 2008; 4(3): 225–233.
  19. 19. Van de Waal DB, John U, Ziveri P, Reichart G-J, Hoins M, Sluijs A, et al. Ocean Acidification Reduces Growth and Calcification in a Marine Dinoflagellate. Plos one. 2013; 8 (6): e65987. pmid:23776586
  20. 20. Mercado J M, Javier F, Gordillo L, Niell F X, Figueroa F L. Effects of different levels of CO2 on photosynthesis and cell components of the red alga Porphyra leucosticta. 1999; J Appl Phycol 11(5): 455–461.
  21. 21. Ihnken S, Roberts S, Beardall J. Differential responses of growth and photosynthesis in the marine diatom Chaetoceros muelleri to CO2 and light availability. Phycologia. 2011; 50(2): 182–193
  22. 22. Gao K, Helbling EW, Häder DP, Hutchins DA. Ocean acidification and marine primary producers under the sun: interactions between CO2, warming, and solar radiation. Mar. Ecol. Prog. Ser. 2012; 470: 167–189.
  23. 23. Torstensson A, Chierici M, Wulff A. The influence of increased temperature and carbon dioxide levels on the benthic/sea ice diatom Navicula directa. Polar Biol. 2012; 35(2): 205–214
  24. 24. Flynn KJ, Blackford JC, Baird ME, Raven JA, Clark DR. Beardall J, et al. Changes in pH at the exterior surface of plankton with ocean acidification. Nature climate change. 2012;2(7): 510–513
  25. 25. Yang Z B, Hodgkiss I J. Hong Kong’s worst “red tide”—causative factors reflected in a phytoplankton study at Port Shelter station in 1998. Harmful Algae. 2004; 3(2):149–161.
  26. 26. Tangen K. Blooms of Gyrodinium aureolum (Dinophygeae) in North European waters, accompanied by mortality in marine organisms. Sarsia. 1977; 63(2): 123–133.
  27. 27. Vanhoutte-Brunier A, Fernand L, Ménesguen A, Lyons S, Gohin F, Cugier P. Modelling the Karenia mikimotoi bloom that occurred in the western English Channel during summer 2003. Ecol Model. 2008; 210(4): 351–376.
  28. 28. Raine R, McMahon T. Physical dynamics on the continental shelf off southwestern Ireland and their influence on coastal phytoplankton blooms. Cont Shelf Res. 1998; 18(8): 883–914.
  29. 29. Godhe A, Otta S K, Rehnstam-Holm A S, Karunasagar I, Karunasagar I. Polymerase Chain Reaction in Detection of Gymnodinium mikimotoi and Alexandrium minutum in Field Samples from Southwest India. Mar Biotechnol. 2001; 3(2):152–62. pmid:14961378
  30. 30. Yao W, Li C, Gao J. Red tide plankton along the south coastal area in Zhejiang province. Marine Science Bulletin. 2006; 25(3):87–91.
  31. 31. Guillard R R L. Culture of phytoplankton for feeding marine invertebrates. In:Culture of marine invertebrate animals. Springer.1975; 29–60.
  32. 32. Fu FX, Warner ME, Zhang Y, Feng Y, Hutchins DA. Effects of increased temperature and CO2 on photosynthesis, growth, and elemental ratios in marine Synechococcus and Prochlorococcus (Cyanobacteria) J Phycol. 2007;43(3): 485–496.
  33. 33. Fu F X, Zhang Y, Warner M E, Feng Y, Sun J, Hutchins DA. A comparison of future increased CO2 and temperature effects on sympatric Heterosigma akashiwo and Prorocentrum minimum. Harmful Algae. 2008;7(1): 76–90.
  34. 34. Hutchins D A, Fu F X, Zhang Y, Warner M E, Feng Y, Portune K, et al. CO2 control of Trichodesmium N2 fixation, photosynthesis, growth rates, and elemental ratios: Implications for past, present, and future ocean biogeochemistry. Limnology and Oceanography.2007;52(4): 1293–1304.
  35. 35. Lewis E, Wallace D, Allison LJ. Program developed for CO2 system calculations. Tennessee: Carbon Dioxide Information Analysis Center, managed by Lockheed Martin Energy Research Corporation for the US Department of Energy,1998.
  36. 36. Zhao Y, Wang Y, Quigg A. The 24 hour recovery kinetics from N starvation in Phaeodactylum tricornutum and Emiliania huxleyi. J phycol. 2015; 51(4): 726–738. pmid:26986793
  37. 37. Fourqurean J W, Zieman J C, Powell G V N. Phosphorus limitation of primary production in Florida Bay: evidence from C: N: P ratios of the dominant seagrass Thalassia testudinum. Limnol Oceanogr. 1992; 37(1): 162–171.
  38. 38. Porra R J. The chequered history of the development and use of simultaneous equations for the accurate determination of chlorophylls a and b. Photosynth Res. 2002; 73: 149–156. pmid:16245116
  39. 39. Platt T, Gallegos C L, Harrison W G. Photoinhibition of photosynthesis in natural assemblages of marine phytoplankton. J Mar Res. 1980; 38: 687–701.
  40. 40. Ralph PJ, Gademann R. Rapid light curves: a powerful tool to assess photosynthetic activity. Aquat Bot. 2005; 82(3): 222–237.
  41. 41. Gerard V A, Driscoll T. A spectrophotometric assay for rubisco activity: application to the kelp laminaria saccharina and implications for radiometric assays1. J phycol. 1996; 32(5): 880–884
  42. 42. Wilbur K M, Anderson N G. Electrometric and colorimetric determination of carbonic anhydrase. J Biol Chem. 1948; 176(1): 147–154. pmid:18886152
  43. 43. Moroney JV, Husic HD, Tolbert NE. Effect of carbonic anhydrase inhibitors on inorganic carbon accumulation by Chlamydomonas reinhardtii. Plant Physiol. 1985;79(1): 177–183. pmid:16664365
  44. 44. Axelsson L, Ryberg H, Beer S. Two modes of bicarbonate utilization in the marine green macroalga Ulva lactuca. Plant Cell Environ. 1995;18(4): 439–445.
  45. 45. Raven J A, Johnston A M. Mechanisms of inorganic-carbon acquisition in marine phytoplankton and their implications for the use of other resources. Limnology and Oceanography. 1991; 36(8):1701–1714.
  46. 46. Maberly SC. Exogenous sources of inorganic carbon for photosynthesis by marine macroalgae 1. J Phycol. 1999; 26(3): 439–449.
  47. 47. Johnston A M, Maberly S C, Raven J A. The acquisition of inorganic carbon by four red macroalgae. Oecologia.1992; 92(3): 317–326. pmid:28312597
  48. 48. Rost B, Riebesell U, Burkhardt S, Sültemeyer D. Carbon Acquisition of Bloom-Forming Marine Phytoplankton. Limnol and Oceanogr. 2003; 48(1):55–67.
  49. 49. Mercado J M, Niell F X. Carbonic anhydrase activity and use of HCO3- in Bostrychia scorpioides (Ceramiales, Rhodophyceae). Eur J Phycol. 1999; 34:13–19
  50. 50. Rost B, Richter K U, Riebesell U, Hansen P J. Inorganic carbon acquisition in red tide dinoflagellates. Plant Cell Environ. 2006; 29(5): 810–822. pmid:17087465
  51. 51. Eberlein T, Van de Waal DB, Rost B. Differential effects of ocean acidification on carbon acquisition in two bloom-forming dinoflagellate species. Physiol plantarum. 2014;151(4): 468–479.
  52. 52. Drechsler Z, Sharkia R, Cabantchik Z I, Beer S. Bicarbonate uptake in the marine maxroalga Ulva sp. is inhibited by classical probes of anion exchange by red blood cells. Planta. 1993; 191:34–40
  53. 53. Larsson C, Axelsson L, Ryberg H, Beer S. Photosynthetic carbon utilization by Enteromorpha intestinalis (Chorophyta) form a Swedish rockpool. Eur J Phycol. 1997; 32:49–54
  54. 54. Trimborn S, Lundholm N, Thoms S, Richter K U, Krock B, Hansen P J, et al. Inorganic carbon acquisition in potentially toxic and non-toxic diatoms: the effect of pH-induced changes in seawater carbonate chemistry. Physiol Plantarum. 2008; 133(1):92–105.
  55. 55. Trimborn S, Wolf-Gladrow D, Richter K U, Rost B. The effect of pCO2, on carbon acquisition and intracellular assimilation in four marine diatoms. J Exp Mar Biol Ecol. 2009; 376(1):26–36.
  56. 56. LOW-DÉCARIE E, Fussmann GF, Bell G. The effect of elevated CO2 on growth and competition in experimental phytoplankton communities. Global Change Biology. 2011;17(8): 2525–2535.58.
  57. 57. Gordillo F J L, Niell F X, Figueroa F L. Non-photosyntheticenhancement of growth by high CO2 level in the nitrophilic seaweed Ulva rigida C. Agardh (Chlorophyta). Planta.2001;213:64–70 pmid:11523657
  58. 58. Chen X, Gao K. Effect of CO2 concentrations on the activity of photosynthetic CO2 fixation and extracelluar carbonic anhydrase in the marine diatom Skeletonema costatum. Chinese Sci Bull. 2003;48(23): 2616–2620.
  59. 59. Nielsen L T, Hallegraeff G M, Wright S W, Hansen P J. Effects of experimental seawater acidification on an estuarine plankton community. Aquat Microb Ecol. 2011; 65(3): 271–285.
  60. 60. Gao K, Ruan Z, Villafane V E, Gattuso J P, Helbling E W. Ocean acidification exacerbates the effect of UV radiation on the calcifying phytoplankter Emiliania huxleyi. Limnol Oceanogr. 2009; 54(6): 1855–1862.
  61. 61. Montechiaro F, Giordano M. Compositional homeostasis of the dinoflagellate Protoceratium reticulatum grown at three different pCO2. J Plant Physiol. 2010; 167(2): 110–113. pmid:19740567
  62. 62. Beardall J, Giordano M. Ecological implications of microalgal and cyanobacterial CO2 concentrating mechanisms, and their regulation. Funct Plant Biol. 2002;29(3): 335–347.
  63. 63. Mercado J M, Javier F, Gordillo L, Niell F X, Figueroa F L. Effects of different levels of CO2 on photosynthesis and cell components of the red alga Porphyra leucosticta. J Appl Phycol. 1999; 11(5):455–461.
  64. 64. Gordillo F J L, Niell F X, Figueroa F L. Non-photosynthetic enhancement of growth by high CO2 level in the nitrophilic seaweed Ulva rigida C. Agardh (Chlorophyta). Planta. 2001; 213(1):64–70. pmid:11523657
  65. 65. Hennon G M M, Quay P, Morales R, Swanson L M, Armbrust E V. Acclimation conditions modify physiological response of the diatom Thalassiosira pseudonana to elevated CO2 concentrations in a nitrate-limited chemostat. J phycol. 2014; 50(2): 243–253. pmid:26988182
  66. 66. Geider R, Osborne B A. Respiration and microalgal growth: a review of the quantitative relationship between dark respiration and growth. New Phytologist. 1989; 12(3): 327–341.
  67. 67. Amthor J S. Respiration in a future, higher CO2 world. Plant Cell Environ. 1991; 14:13–20
  68. 68. Falkowski P G, Raven J A. Aquatic photosynthesis. Princeton University Press. 2013.
  69. 69. Li Q, Canvin DT. Energy Sources for HCO3− and CO2 Transport in Air-Grown Cells of Synechococcus UTEX 625. Plant physiol. 1998;116(3): 1125–1132. pmid:9501145
  70. 70. Tchernov D, Helman Y, Keren N, Luz B, Ohad I, Reinhold L, et al. Passive entry of CO2 and its energy-dependent intracellular conversion to HCO3− in Cyanobacteria are driven by a photosystem I-generated ΔμH+. Biol Chem. 2001; 276(26): 23450–23455.
  71. 71. Gao K, Xu J, Gao G, et al. Rising CO2 and increased light exposure synergistically reduce marine primary productivity. Nature Climate Change. 2012; 2(7):519–523.