Next Article in Journal
Antihyperglycemic and Antilipidemic Properties of a Tea Infusion of the Leaves from Annona cherimola Miller on Streptozocin-Induced Type 2 Diabetic Mice
Next Article in Special Issue
Chemical Composition and Potential Practical Application of 15 Red Algal Species from the White Sea Coast (the Arctic Ocean)
Previous Article in Journal
A Spectroscopic Overview of Intramolecular Hydrogen Bonds of NH…O,S,N Type
Previous Article in Special Issue
Cyanobacteria as Natural Therapeutics and Pharmaceutical Potential: Role in Antitumor Activity and as Nanovectors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Metabolomic Study of Heterotrophically Grown Chlorella sp. Isolated from Wastewater in Northern Sweden

by
Jean Claude Nzayisenga
and
Anita Sellstedt
*
Department of Plant Physiology, UPSC, Umeå University, 901 87 Umeå, Sweden
*
Author to whom correspondence should be addressed.
Molecules 2021, 26(9), 2410; https://doi.org/10.3390/molecules26092410
Submission received: 22 March 2021 / Revised: 15 April 2021 / Accepted: 16 April 2021 / Published: 21 April 2021

Abstract

:
There are numerous strains of Chlorella with a corresponding variety of metabolic pathways. A strain we previously isolated from wastewater in northern Sweden can grow heterotrophically as well as autotrophically in light and has higher lipid contents under heterotrophic growth conditions. The aims of the present study were to characterize metabolic changes associated with the higher lipid contents in order to enhance our understanding of lipid production in microalgae and potentially identify new compounds with utility in sustainable development. Inter alia, the amino acids glutamine and lysine were 7-fold more abundant under heterotrophic conditions, the key metabolic intermediate alpha-ketoglutarate was more abundant under heterotrophic conditions with glucose, and maltose was more abundant under heterotrophic conditions with glycerol than under autotrophic conditions. The metabolite 3-hydroxy-butyric acid, the direct precursor of the biodegradable plastic PHB (poly-3-hydroxy-butyric acid), was also more abundant under heterotrophic conditions. Our metabolomic analysis has provided new insights into the alga’s lipid production pathways and identified metabolites with potential use in sustainable development, such as the production of renewable, biodegradable plastics, cosmetics, and nutraceuticals, with reduced pollution and improvements in both ecological and human health.

1. Introduction

Microalgae can play important roles in sustainable development and amelioration of climate change [1,2], both as potential sources of renewable fuel (biodiesel) and substances such as the biodegradable plastic polyhydroxy butyric acid (PHB, from poly-3-hydroxy-butyric acid). There is a direct comparison that can be drawn between micro particles from plastic and diesel particles. Diesel engines exhaust small particles that are less than 2.5 μm [3]. These particles are capable of crossing cell membranes and triggering oxidative stress and inflammation that have been linked to increased risk of cardiovascular and respiratory diseases and lung cancer [4]. Taking this into account, it is urgent to decrease the use of plastics from oil-based materials by increasing the use of products from microalgae.
Microalgae frequently live in varying environments of light, salinity, and temperature. Thus, the production of algal-derived, biologically active compounds may be determined by the selection of appropriate cultivation conditions, making these algae natural bioreactors. Cultivation methods were reviewed [5], and it was found that the carbon assimilation and energy dissipation pathways strongly differ between autotrophically, mixotrophically, and heterotrophically grown algae. Thus, for example, the heterotrophic growth on glucose of C. vulgaris is reportedly superior to both autotrophic and mixotrophic cultivation for such purposes [6], with particularly high productivity of biodiesel and other biorefinery products [5]. We previously identified a strain of Chlorella that had a high lipid content (39.5%), good biodiesel quality, and low carbohydrate contents (8.2%) when grown under heterotrophic conditions with glycerol as the carbon source. However, the alga’s carbohydrate contents were higher under autotrophic and mixotrophic conditions; thus, an appropriate modulation of growth conditions could be used to enhance the production of bioethanol from its biomass [7]. A strain of C. zofingiensis reportedly has higher contents of total lipids, neutral lipids, triacylglycerols (TAGs, key precursors of biodiesel production), and oleic acid under heterotrophic conditions [8]. In stark contrast, ref. [9] found that the lipid content of the yellow-green alga Tribonema minus was much higher under autotrophic conditions than under heterotrophic conditions. Of course, diverse variables influence lipid production, and attempts to improve the financial viability of biofuel production from some microalgae have been shown. Inter alia, nitrogen starvation can substantially increase their TAG generation; thus, this is the most common approach for enhancing their lipid production [10]. The concentration of one precursor of biodiesel production, namely triacylglycerol, was shown to significantly increase in certain microalgae species upon nitrogen starvation [10]. All these studies described different aspects of algae under different growth conditions. However, the understanding of these variations is currently hindered by a paucity of detailed metabolomic information on heterotrophically grown algae [11], relative to the vast amounts of metabolomic information available on autotrophically and mixotrophically grown algae [12,13]. This prompted us to compare metabolomic profiles of the Chlorella strain we previously isolated under heterotrophic and autotroph growth conditions. We also compared cultures of the alga grown with both glucose and glycerol as the sole carbon source, as the former is widely used for this purpose, but the latter is cheaper [14,15]. Glycerol is absorbed by algae that can use it as a carbon source via simple diffusion, then transformed by reaction with ATP into glyceraldehyde-3-phosphate, an intermediate in glycolysis, and subsequently triose phosphate. The mechanisms and enzymes involved are well understood [5], but the mechanisms involved in the higher lipid production under heterotrophic conditions with glycerol as the carbon source remain unclear. Thus, the main aims of the study presented here were to characterize the metabolic changes associated with the alga’s higher lipid contents under heterotrophic growth conditions. Such information is important for modulating lipid production in microalgae and hence the production of diverse substances with potential utility in sustainable development, with benefits for both ecological and human health.

2. Materials and Methods

2.1. Growth Assessment

To examine the effects of the growth conditions on the alga’s biomass, sets of eight cultures of the locally isolated microalga strain Chlorella sp. [7] were grown in autotrophic, mixotrophic, and heterotrophic conditions for 8 days. In all cases, the basal growth medium was municipal wastewater collected and immediately stored at 4 °C for at most 10 days. Thereafter, the wastewater was filtered, autoclaved, and otherwise treated as previously described [7]. Algal inoculum was grown in 100 mL portions of BG11 in 500 mL flasks under autotrophic conditions (16 h light/8 h dark cycles, with 120 μmol/m2 s−1 illumination during the light phase and constant bubbling with 5% CO2, at 25 °C). To start each growth experiment, 200 mL of wastewater in a 1 l flask was inoculated with inoculum to an optical density (OD630) of 0.06 (equivalent to 0.005 g algal biomass). In autotrophic and mixotrophic conditions, cultures were bubbled with 5% CO2, and both mixotrophic and heterotrophic cultures were supplemented with either glycerol or glucose (VWR International, Pennsylvania, PA, USA) to a final concentration of 37.5 mM.

2.2. Harvesting

The biomass was collected after 8 days of growth, and after washing twice with distilled water, the cells were recovered by centrifugation (3700× g, 6 min), immediately placed in liquid N2, then freeze-dried for 3 days and weighed.

2.3. Sample Preparation for Metabolomic Analysis

Samples were prepared as previously described [16]. Briefly, 1 mL of extraction buffer (20:20:60 v/v chloroform:water:methanol) including internal standards was added to 9–12 mg of each sample. The samples were shaken with a tungsten bead in a mixer mill at 30 Hz for 3 min, the bead was removed, and the sample was centrifuged at +4 °C, 14,000 rpm, for 10 min. Next, 200 µL of supernatant was transferred to a micro vial, and solvents were evaporated.

2.4. Derivatization

Derivatization was performed as earlier described [16]. Briefly, 30 µL of methoxyamine (15 µg/µL in pyridine) were added to the dry sample, which was shaken vigorously for 10 min before left to react at room temperature for 16 h. Following addition of 30 µL of N-methyl-N-(trimethylsilyl)trifluoroacetamide (MSTFA), the samples were shaken and incubated for a further hour at room temperature. A 30 µL portion of methyl stearate solution (15 ng/µL in heptane) was subsequently added, and the samples were subjected to gas chromatography-mass spectrometry (GC-MS, Agilent 7890, Santa Clara, CA, USA) analysis, as previously described [16,17]. Briefly, 1 µL of the derivatized sample was injected in either splitless or split (1:20) mode by an CTC Combi Pal Xt Duo (CTC Analytics AG, Zwingen, Switzerland) autosampler into an Agilent 7890A gas chromatograph equipped with a 30 m × 0.25 mm i.d. fused-silica capillary column with a chemically bonded 0.25 μm DB 5-MS UI stationary phase (J&W Scientific, Folsom, CA, USA). The voltage was turned on after a 290 s solvent delay.

2.5. Data Evaluation

Raw MS files from the metabolic analyses were exported from ChromaTOF software (LECO, St Joseph, MI, USA) in NetCDF format to Matlab™ (Mathworks, Natick, MA, USA). Following pre-treatment procedures, such as baseline correction and chromatogram alignment in Matlab, the data were further processed by peak integration and multivariate curve resolution using custom scripts. Next, analytes were identified by comparison of their retention time indices and mass spectra with those of entries in both in-house and public databases.
Finally, differences between cultures grown under different conditions were explored by principal coordinate analysis (PCA) using SIMCA (Umetrics, Umeå, Sweden). The differences between algae grown heterotrophically with glycerol, heterotrophically with glycerol, autotrophically, and mixotrophically were analysed by t-tests.

3. Results and Discussion

Algal heterotrophic metabolism can occur in the light as well as the dark [18]. However, the metabolic pathways involved differ and require elucidation in order to optimize algal exploitation for production of biodiesel and/or other substances. For example, reduced power and energy (ATP) are obtained via glycolysis, the pentose phosphate and tricarboxylic acid pathways linked to mitochondrial electron transport pathways, in dark, aerobic conditions, but via photosynthesis in autotrophic conditions [18]. This study provides more detailed information on the metabolic profiles, and hence key pathways, of the local Chlorella strain under four conditions: autotrophic, mixotrophic, and heterotrophic with glycerol or glucose as the sole carbon source.

3.1. Growth of Chlorella Vulgaris

After 8 days of growth, there was no significant difference in biomass between cultures grown autotrophically, mixotrophically with glycerol, and heterotrophically with glucose (Table 1). However, cells grown mixotrophically with glucose and heterotrophically with glycerol had accumulated less biomass than autotrophic cells (Table 1). These results are consistent with previous findings regarding the growth of Chlorella sp. and the amounts of phosphorous and nitrogen taken up by heterotrophic grown cultures [7]. It was shown that the amount of phosphorous and nitrogen taken up by heterotrophically grown cultures with glycerol were 5 and 19 mg·L−1, corresponding to 83 and 80% uptake, respectively [7].

3.2. Metabolomic Analysis of Chlorella Alga Under Auto-, Mixo-, and Heterotrophic Conditions

The GC-MS analysis identified 74 distinct metabolites that were present in all samples (Table 2). Key metabolites and metabolite classes separating the samples are shown in the constructed traditional principal coordinate analysis PCA plot presented in Figure 1, which clearly reveals correlations between the abundance of metabolites and growth conditions. It shows that the growth conditions substantially influenced the levels of certain metabolites, inter alia there were major metabolomic differences between cells grown heterotrophically with glucose and glycerol; thus, it was clearly shown that the samples could be separated based on the growth conditions. Moreover, there were major differences between cells grown heterotrophically in glucose and glycerol (Figure 1).
The relative abundances of selected metabolites presented in Figure 2, Figure 3, Figure 4 and Figure 5 clearly show cultivation-condition-related metabolomic variations. The levels of amino acids (notably alanine, ornithine, and aspartic acid) were several times higher following heterotrophic growth with glycerol than under other growth conditions (Figure 2). This has strong physiological (as well as metabolomic) implications because in addition to being building blocks of proteins, amino acids are precursors of diverse N-containing molecules such as nucleic acids, polyamines, quaternary ammonium compounds, and some hormones.
Under environmental stress, de novo protein synthesis is generally inhibited, and protein turnover and proteolytic activity are increased, resulting in increases in the total free amino acid content [9,19]. In addition, ref. [19] also found that the N/C assimilation and distribution pathways related to the glutamate-glutamine system, amino acid (GABA) catabolism and synthesis, and the TCA cycle and glycolysis contribute to the shunting of excess carbon into lipid biosynthesis.
To extend our understanding of the links between growth conditions and metabolite contents in Chlorella sp., we focused particularly on three metabolites that play important roles in carbon metabolism: glucose-6-phosphate, fructose-6-phosphate, and glyceric acid-3-phosphate. All three compounds were much less abundant in cells grown heterotrophically with glycerol than in cells grown autotrophically (Table 2 and Figure 3). In addition, fructose-6-phosphate was the most abundant carbohydrate and much more abundant under autotrophic conditions than under the other conditions (Table 2 and Figure 3). As the formation of fructose-6-phosphate is the rate-limiting step in gluconeogenesis, the low levels of glucose-6-phosphate, fructose-6- phosphate, and glyceric acid-3-phosphate observed under heterotrophy with glycerol as the sole carbon source may be associated with low gluconeogenesis rates. This would explain the alga’s high carbohydrate contents among the microalgae under autotrophic conditions (Figure 3). Fructose-1,6-bisphosphatase, which catalyzes the conversion of fructose-1,6-bisphosphate into fructose-6-phosphate during gluconeogenesis, is also reportedly up-regulated in microalgae with low lipid content and high carbohydrate content [9,20]. This corroborates the indication that the higher carbohydrate contents observed under all the conditions except heterotrophy with glycerol are associated with higher contents of sugars involved in glucose synthesis.
Interestingly, several fatty acids (e.g., eicosanoid, hetadecanoic, nonanoic, arachidic, and octadecanoic acids) and the total fatty acid content were more abundant under heterotrophic conditions with glycerol than under autotrophic conditions (Table 2; Figure 4). These results support our previous finding of a correlation between heterotrophic conditions with glycerol as the carbon source and high lipid production [7]. We have also recently observed an association between high linolenic acid (18:3) content and low oleic acid (18:1) content under heterotrophic conditions with glycerol, which improves biodiesel quality [7]. Increased levels of fatty acids such as nonanoic acid (17:0) and arachidic acid (20:0) (Table 2) support our previous results that heterotrophic conditions with glycerol as the carbon source are associated with increased lipid production [7].
A metabolite of particular interest that we detected is 3-hydroxy-butyric acid (Table 2; Figure 5), which is generally obtained from renewable carbohydrates by fermentation [1,2]. Together with component polyhydroxy-alkanoates (PHAs) and starch-, cellulose-, and protein-based substances, this is a major candidate for commercial bioplastic production [21,22]. Moreover, we have previously shown that the local Chlorella sp. can produce more lipids (including 18:1 fatty acid), which is important for biodiesel quality [7]. Similarly, [13] found that illumination significantly affected the fatty acid saturation level in the alga Euglena gracilis, and that while it could accumulate considerable amounts of short-chain fatty acids in the dark, its desaturase activity (and hence unsaturated fatty acid content) was higher in the light.

3.3. Bioplastics

Commonly used plastics originating from oil-based materials are estimated to reach 622 MT by year 2034 [23]. Microplastics are either produced by design or are formed as a result of degradation of macroplastics. An increasing amount of evidence suggests a widespread exposure to microplastics from foods, drinking water, and air [3]. Replacing microplastics with bioplastics produced from algae would be beneficial for human health. Despite a rapid increase in the bioplastic production, it was still only 2.11 MT in 2018 [23]. The component 3-hydroxy-butyric acid is, together with polyhydroxy-alkanoates (PHAs) and starch-, cellulose-, and protein-based substances, a major candidate for commercial bioplastic production [21,22]. It was found in higher amounts under heterotrophic conditions with glycerol in this study (Figure 5).

4. Conclusions

In this study, we have identified key metabolites connected to the high lipid contents of a Chlorella strain (isolated in northern Sweden) when grown under heterotrophic conditions with glycerol as the sole carbon source. Under all the other growth conditions, we observed higher contents of glucose-6-phosphate, fructose-6-phosphate, and glyceric acid-3-phosphate, which may be associated with higher gluconeogenesis rates. Growth under heterotrophic conditions with glycerol resulted in the highest levels of certain amino acids, for reasons that are unclear and warrant further attention. Moreover, levels of 3-hydroxybutyric acid, which can be potentially used for bioplastic production, were highest under heterotrophic conditions. Thus, these findings may enhance our understanding of lipid production in microalgae and facilitate their biorefinery-based exploitation, with substantial ecological and human health benefits.

Author Contributions

A.S. and J.C.N. formulated the experiment. J.C.N. performed the experiment. A.S. was responsible for supervision and writing of the first version of the manuscript. J.C.N. and A.S. contributed equally to the last version of the manuscript and have read and agreed to the published version of the manuscript.

Funding

Financial support for the project were provided by Energimyndigheten, Sweden (project ref. Microbiorefine 2013-06259) and FORMAS, Sweden (project ref. 942-2015-92).

Data Availability Statement

Not applicable.

Acknowledgments

We thank A. Johansson (A.J.) and H. Stenlund (H.S.) at the Swedish Metabolomic Centre for their skillful analysis.

Conflicts of Interest

The authors declare no conflict of financial interests.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Abdo, S.M.; Ali, G.H. Analysis of polyhydroxybutrate and bioplastic production from microalgae. Bull. Natl. Res. Cent. 2019, 43, 1–4. [Google Scholar] [CrossRef]
  2. Zhang, C.; Show, P.-L.; Ho, S.-H. Progress and perspective on algal plastics—A critical review. Bioresour. Technol. 2019, 289, 121700. [Google Scholar] [CrossRef] [PubMed]
  3. Vethaak, A.D.; Legler, J. Microplastics and human health. Science 2021, 371, 672–674. [Google Scholar] [CrossRef] [PubMed]
  4. Kelly, F.J.; Mudway, I.S.; Fussell, J.C. Air Pollution and Asthma: Critical Targets for Effective Action. Pulm. Ther. 2020, 1–16. [Google Scholar] [CrossRef]
  5. Perez-Garcia, O.; Bashan, Y. Microalgal Heterotrophic and Mixotrophic Culturing for Bio-refining: From Metabolic Routes to Techno-economics. In Algal Biorefineries: Volume 2: Products and Refin-ery Design; Prokop, A., Bajpai, R.K., Zappi, M.E., Eds.; Springer International Publishing: Cham, Switzerland, 2015; Volume 2, pp. 61–131. [Google Scholar]
  6. Perez-Garcia, O.; De-Bashan, L.E.; Hernandez, J.-P.; Bashan, Y. Efficient of growth and nutrient uptake from wastewater by heterotrophic, autotrophic and mixotrophic cultivation of chlorella vul-garis immobilized with Azospirillum blasiliense. J. Phycol. 2010, 46, 800–812. [Google Scholar] [CrossRef]
  7. Nzayisenga, J.C.; Eriksson, K.; Sellstedt, A. Mixotrophic and heterotrophic production of lipids and carbohydrates by a locally isolated microalga using wastewater as a growth medium. Bioresour. Technol. 2018, 257, 260–265. [Google Scholar] [CrossRef]
  8. Liu, J.; Huang, J.; Sun, Z.; Zhong, Y.; Jiang, Y.; Chen, F. Differential lipid and fatty acid profiles of photoautotrophic and heterotrophic Chlorella zofingiensis: Assessment of algal oils for biodiesel pro-duction. Bioresour. Technol. 2011, 102, 106–110. [Google Scholar] [CrossRef]
  9. Wang, H.; Zhang, Y.; Zhou, W.; Noppol, L.; Liu, T. Mechanism and enhancement of lipid accumu-lation in filamentous oleaginous microalgae Tribonema minus under heterotrophic condition. Biotechnol. Biofuels 2018, 1, 328. [Google Scholar] [CrossRef] [Green Version]
  10. Zienkiewicz, K.; Du, Z.-Y.; Ma, W.; Vollheyde, K.; Benning, C. Stress-induced neutral lipid bio-synthesis in microalgae—Molecular, cellular and physiological insights. Biochim. Biophys. Acta 2016, 186, 1269–1281. [Google Scholar] [CrossRef] [Green Version]
  11. Morales-Sánchez, D.Z.; Martinez-Rodriguez, O.A.; Kyndt, J.; Martinez, A. Heterotrophic growth of microalgae: Metabolic aspects. World J. Microbiol. Biotechnol. 2015, 31, 1–9. [Google Scholar] [CrossRef]
  12. Renberg, L.; Johansson, A.I.; Shutova, T.; Stenlund, H.; Aksmann, A.; Raven, J.A.; Gardeström, P.; Moritz, T.; Samuelsson, G. A Metabolomic Approach to Study Major Metabolite Changes during Acclimation to Limiting CO2 in Chlamydomonas reinhardtii. Plant Physiol. 2010, 154, 187–196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Zeng, M.; Hao, W.; Zou, Y.; Shi, M.; Jiang, Y.; Xiao, P.; Lei, A.; Hu, Z.; Zhang, W.; Zhao, L.; et al. Fatty acid and metabolomic profiling approaches differentiate heterotrophic and mixo-trophic culture conditions in a microalgal food supplement ‘Euglena’. BMC Biotechnol. 2016, 16, 49. [Google Scholar] [CrossRef] [Green Version]
  14. Li, X.; Xu, H.; Wu, Q. Large-scale biodiesel production from microalgaChlorella protothecoides through heterotrophic cultivation in bioreactors. Biotechnol. Bioeng. 2007, 98, 764–771. [Google Scholar] [CrossRef]
  15. Vivek, N.; Sindhu, R.; Madhavan, A.; Anju, A.J.; Castro, E.; Faraco, V.; Pandey, A.; Binod, P. Recent advances in the production of value added chemicals and lipids utilizing biodiesel industry generated crude glycerol as a substrate—Metabolic aspects, challenges and possibilities: An overview. Bioresour. Technol. 2017, 239, 507–517. [Google Scholar] [CrossRef] [PubMed]
  16. Gullberg, J.; Jonsson, P.; Nordstrom, A.; Sjostrom, M.; Moritz, T. Design of experiments: An effi-cient strategy to identify factors influencing extraction and derivatization of Arabidopsis thaliana samples in metabolomic studies with gas chromatography/mass spectrometry. Anal. Biochem. 2004, 331, 283–295. [Google Scholar] [CrossRef] [PubMed]
  17. Jiye, A.; Trygg, J.; Gullberg, J.; Johansson, A.I.; Jonsson, P.; Antti, H.; Marklund, S.L.; Moritz, T. Ex-traction and GC/MS Analysis of the Human Blood Plasma Metabolome. Anal. Chem. 2005, 77, 8086–8094. [Google Scholar]
  18. Mohan, S.V.; Rohit, M.V.; Chiranjeevi, P.; Chandra, R.; Beeramganti, N. Heterotrophic Mi-croalgae Cultivation to Synergize Biodiesel Production with Waste Remediation: Progress and Per-spectives. Bioresour. Technol. 2015, 184, 169–178. [Google Scholar] [CrossRef] [PubMed]
  19. Chen, H.; Zheng, Y.; Zhan, J.; He, C.; Wang, Q. Comparative metabolic profiling of the li-pid-producing green microalga Chlorella reveals that nitrogen and carbon metabolic pathways con-tribute to lipid metabolism. Biotechnol. Biofuels 2017, 10, 153. [Google Scholar] [CrossRef] [Green Version]
  20. Yang, Z.-K.; Niu, Y.-F.; Ma, Y.-H.; Xue, J.; Zhang, M.-H.; Yang, W.; Liu, J.-S.; Lu, S.; Guan, Y.; Li, H.-Y. Molecular and cellular mechanisms of neutral lipid accumulation in diatom following ni-trogen deprivation. Biotechnol. Biofuels 2013, 6, 67. [Google Scholar] [CrossRef] [Green Version]
  21. Mekonnen, T.; Mussone, P.G.; Khalil, H.; Bressler, D.C. Progress in bio-based plastics and plasticizing modifications. J. Mater. Chem. A 2013, 1, 13379–13398. [Google Scholar] [CrossRef] [Green Version]
  22. Zheng, J.; Suh, S. Strategies to reduce the global carbon footprint of plastics. Nat. Clim. Chang. 2019, 9, 374–378. [Google Scholar] [CrossRef]
  23. Lamberti, F.M.; Román-Ramírez, L.A.; Wood, J. Recycling of Bioplastics: Routes and Benefits. J. Polym. Environ. 2020, 28, 2551–2571. [Google Scholar] [CrossRef]
Figure 1. PCA score plot showing metabolomic separation of sets of eight replicates of Chlorella cultures after 8 days of cultivation in the following growth conditions: autotrophic (A), mixotrophic with glycerol (YM), heterotrophic with glycerol (YH), mixotrophic with glucose (UM), heterotrophic with glucose (UH). Eight replicates for each growth condition are shown.
Figure 1. PCA score plot showing metabolomic separation of sets of eight replicates of Chlorella cultures after 8 days of cultivation in the following growth conditions: autotrophic (A), mixotrophic with glycerol (YM), heterotrophic with glycerol (YH), mixotrophic with glucose (UM), heterotrophic with glucose (UH). Eight replicates for each growth condition are shown.
Molecules 26 02410 g001
Figure 2. Relative contents (means ± standard deviations, n = 8) of three amino acids in the Chlorella cultures after 8 days of cultivation in the following growth conditions: autotrophic (AUTO), mixotrophic with glycerol (GLY MIXO), heterotrophic with glycerol (GLY HETE), mixotrophic with glucose (GLU MIXO), heterotrophic with glucose (GLU HETE). *, **, and *** indicate significant differences in concentration relative to cultures grown in heterotrophic conditions with glycerol according to Student’s t-test at p < 0.05, 0.01, and 0.001, respectively.
Figure 2. Relative contents (means ± standard deviations, n = 8) of three amino acids in the Chlorella cultures after 8 days of cultivation in the following growth conditions: autotrophic (AUTO), mixotrophic with glycerol (GLY MIXO), heterotrophic with glycerol (GLY HETE), mixotrophic with glucose (GLU MIXO), heterotrophic with glucose (GLU HETE). *, **, and *** indicate significant differences in concentration relative to cultures grown in heterotrophic conditions with glycerol according to Student’s t-test at p < 0.05, 0.01, and 0.001, respectively.
Molecules 26 02410 g002
Figure 3. Relative content of three key carbon metabolism intermediates (means ± standard deviations, n = 8) in the Chlorella cultures after 8 days of cultivation in the following conditions: autotrophic (AUTO), mixotrophic with glycerol (GLY MIXO), heterotrophic with glycerol (GLY HETE), mixotrophic with glucose (GLU MIXO), heterotrophic with glucose (GLU HETE). *, **, and *** indicate significant differences in concentration relative to cultures grown in heterotrophic conditions with glycerol according to Student’s t-test at p < 0.05, 0.01, and 0.001, respectively.
Figure 3. Relative content of three key carbon metabolism intermediates (means ± standard deviations, n = 8) in the Chlorella cultures after 8 days of cultivation in the following conditions: autotrophic (AUTO), mixotrophic with glycerol (GLY MIXO), heterotrophic with glycerol (GLY HETE), mixotrophic with glucose (GLU MIXO), heterotrophic with glucose (GLU HETE). *, **, and *** indicate significant differences in concentration relative to cultures grown in heterotrophic conditions with glycerol according to Student’s t-test at p < 0.05, 0.01, and 0.001, respectively.
Molecules 26 02410 g003
Figure 4. Relative contents (means ± standard deviation, n = 8) of three fatty acids in the Chlorella cultures after 8 days of cultivation in the following growth conditions: autotrophic (AUTO), mixotrophic with glycerol (GLY MIXO), heterotrophic with glycerol (GLY HETE), mixotrophic with glucose (GLU MIXO), heterotrophic with glucose (GLU HETE). The data are presented as means ± standard deviations, n = 8. *, **, *** indicate significant differences in metabolite concentration relative to cultures grown in heterotrophic conditions with glycerol according to Student’s t-test, p < 0.05, 0.01, and 0.001, respectively.
Figure 4. Relative contents (means ± standard deviation, n = 8) of three fatty acids in the Chlorella cultures after 8 days of cultivation in the following growth conditions: autotrophic (AUTO), mixotrophic with glycerol (GLY MIXO), heterotrophic with glycerol (GLY HETE), mixotrophic with glucose (GLU MIXO), heterotrophic with glucose (GLU HETE). The data are presented as means ± standard deviations, n = 8. *, **, *** indicate significant differences in metabolite concentration relative to cultures grown in heterotrophic conditions with glycerol according to Student’s t-test, p < 0.05, 0.01, and 0.001, respectively.
Molecules 26 02410 g004
Figure 5. Relative content of 3-hydroxy butyric acid (means ± standard deviations, n = 8) in the Chlorella cultures after 8 days of cultivation in the following conditions: autotrophic (AUTO), mixotrophic with glycerol (GLY MIXO), heterotrophic with glycerol (GLY HETE), mixotrophic with glucose (GLU MIXO), heterotrophic with glucose (GLU HETE). ** and *** indicate significant differences in concentration relative to cultures grown in heterotrophic conditions with glycerol according to Student’s t-test at p < 0.01 and 0.001, respectively.
Figure 5. Relative content of 3-hydroxy butyric acid (means ± standard deviations, n = 8) in the Chlorella cultures after 8 days of cultivation in the following conditions: autotrophic (AUTO), mixotrophic with glycerol (GLY MIXO), heterotrophic with glycerol (GLY HETE), mixotrophic with glucose (GLU MIXO), heterotrophic with glucose (GLU HETE). ** and *** indicate significant differences in concentration relative to cultures grown in heterotrophic conditions with glycerol according to Student’s t-test at p < 0.01 and 0.001, respectively.
Molecules 26 02410 g005
Table 1. Accumulated biomass of the Chlorella strain after 8 days of growth under conditions of cultivation: autotrophic (A), mixotrophic with glycerol (Y M), heterotrophic with glycerol (Y H), mixotrophic with glucose (U M), heterotrophic with glucose (U H). Means ± standard deviations, n = 8.
Table 1. Accumulated biomass of the Chlorella strain after 8 days of growth under conditions of cultivation: autotrophic (A), mixotrophic with glycerol (Y M), heterotrophic with glycerol (Y H), mixotrophic with glucose (U M), heterotrophic with glucose (U H). Means ± standard deviations, n = 8.
Growth Condition
A1.25 ± 0.09
Y M1.31 ± 0.10
Y H0.52 ± 0.03
U H1.29 ± 180
U M0.70 ± 0.05
Table 2. Metabolites detected in Chlorella strain and their retention indexes (RI), compound classes (CC; FA = fatty acids; CA = carboxylic acids; AA = amino acids; EC = energy compound; V = vitamins; P = phenols; ACA = amino carboxylic acid; Carb= carbohydrates; AlA = alkylamines; Carb= carbohydrate; AM = amides; SP = sugar phosphates; L = lipids) and P values. The P values of differences in levels of the metabolites between algae grown in autotrophic and heterotrophic conditions with glycerol are according to t-tests. Mean ± standard errors, n = 8.
Table 2. Metabolites detected in Chlorella strain and their retention indexes (RI), compound classes (CC; FA = fatty acids; CA = carboxylic acids; AA = amino acids; EC = energy compound; V = vitamins; P = phenols; ACA = amino carboxylic acid; Carb= carbohydrates; AlA = alkylamines; Carb= carbohydrate; AM = amides; SP = sugar phosphates; L = lipids) and P values. The P values of differences in levels of the metabolites between algae grown in autotrophic and heterotrophic conditions with glycerol are according to t-tests. Mean ± standard errors, n = 8.
CompoundsRIC.C.p Value
1-myristoylglycerol2382FA<0.0001
2-methylmalic acid1460CA<0.0001
2-oxoisocaproic acid1207FA0.2423
3-hydroxybutyric acid1156CA0.0377
3-hydroxyisobutyric acid1156CA0.0065
4-aminobutyric acid1522CA0.0860
4-hydroxyphenylacetic acid1636CA0.0064
5-hydroxypipecolic acid1588CA0.0139
Adenosine-5-monophosphate3050EC0.0108
Alanine1100AA0.0010
Alpha-ketoglutaric acid1565CA0.0638
Alpha-tocopherol3145V0.5274
Aminomalonic acid1458ACA0.0425
Arabinose/ribose/xylose1672CA<0.0001
Arabitol1697SA0.0002
Asparagine1658AA0.0696
Aspartic acid1507AA0.0011
Benzoic acid1252P0.0420
Beta-alanine1419AA0.0543
Cadaverine1831AlA<0.0001
Citric acid1800CA<0.0001
Dehydroascorbic acid dimer1837V0.0002
Disaccharide2515Carb0.0303
Eicosanoic acid2435FA0.0388
Ethanolamine1260AA0.0003
Fructose-6-phosphate2280SP0.0041
Fumaric acid1342CA0.0112
Glucose1876S0.0393
Glucose-6-phosphate2316/2327SP0.0008
Glutamic acid1607AA0.0485
Glutamine1763AA0.0438
Glyceric acid1318CA<0.0001
Glyceric acid-3-phosphate1787CA<0.0001
Glycerol1267SA<0.0001
Glycerol-2-phosphate1704SP0.0001
Glycerol-3-phosphate1743SP<0.0001
Glycine1120AA0.8101
Glycolic acid1074CA0.0004
Heptadecanoic acid2139L0.0052
Hexadecanoic acid2042L0.0616
Inositol, myo-2074V0.0151
Inositol, scyllo-2008V<0.0001
Inositol-phosphate2396V0.3207
Isoerythritol1479S0.0336
Lactic acid1057CA0.1729
Lysine1848AA0.0232
Lyxose1633S0.0059
Malic acid1474CA0.6389
Maltose/turanose2727S<0.0001
Methylsuccinate1320CA<0.0001
Myristic acid 501844FA0.0027
Nicotinamide1844V0.0002
Nonanoic acid1357FA<0.0001
Octadecatrienoic acid, 6,9,12-(z,z,z)-2191FA0.061
Octadecatrienoic acid, 9,12,15-(z,z,z)-2218FA0.0370
O-phosphoetanolamine1771AM<0.0001
Ornithine1453AA<0.0001
Pantothenic acid1980V0.0043
Phenylalanine1680A0.0036
Putrescine1732AM0.0768
Pyroglutamic acid1518AA0.0315
Pyruvic acid1049CA<0.0001
Raffinose 3373S0.0062
Salicin2488S<0.0001
Succinic acid1308S<0.0001
Sucrose2616S0.2146
Threonine/allo-threonine1372AA0.027205
Trehalose/maltose2721S<0.0001
Tyrosine1933AA0.6952
Valine1208AA0.0455
Xylitol1684S0.0050
Xylulose1654S<0.0001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nzayisenga, J.C.; Sellstedt, A. Metabolomic Study of Heterotrophically Grown Chlorella sp. Isolated from Wastewater in Northern Sweden. Molecules 2021, 26, 2410. https://doi.org/10.3390/molecules26092410

AMA Style

Nzayisenga JC, Sellstedt A. Metabolomic Study of Heterotrophically Grown Chlorella sp. Isolated from Wastewater in Northern Sweden. Molecules. 2021; 26(9):2410. https://doi.org/10.3390/molecules26092410

Chicago/Turabian Style

Nzayisenga, Jean Claude, and Anita Sellstedt. 2021. "Metabolomic Study of Heterotrophically Grown Chlorella sp. Isolated from Wastewater in Northern Sweden" Molecules 26, no. 9: 2410. https://doi.org/10.3390/molecules26092410

Article Metrics

Back to TopTop