Protocol for maximizing the triglycerides-enriched lipids production from Dunaliella salina SA32007 biomass, isolated from the Salar de Atacama (Northern Chile)

Abstract

This paper reports the effects of different culture conditions for Dunaliella salina SA32007 from Salar de Atacama (second Region, northern ofChile) over biomass, lipid production and triglycerides synthesis. A maximum value of microalgae density (8.2 × 109 Cells/L) and an intrinsic growth rate (0.17 d-1), were obtained using a culture with 0.5 mol/L of NaCl and a nitrogen/phosphorous (N/P) limitation of 14/1. The triglycerides production was significantly favoured under nitrogen deficiency conditions (Mann-Whitney test; p = 0.0043). However there was a nitrogen-limiting threshold for the stimulation and accumulation of triglycerides (N/P: 14/1), lower than that limit, these compounds would not be accumulated. It was also observed that triglyceride content was not proportional to the total lipid content and the maximum number of cells. The aeration system employed stimulated the growth and synthesis of structural organic molecules. Regarding microalgae growth stage subjected to nitrogen deficiency, when the deficit was applied before the lag phase, the negative effect on the biomass and the triglycerides production decreased.

Share and Cite:

Arias-Forero, D. , Hayashida, G. , Aranda, M. , Araya, S. , Portilla, T. , García, A. and Díaz-Palma, P. (2013) Protocol for maximizing the triglycerides-enriched lipids production from Dunaliella salina SA32007 biomass, isolated from the Salar de Atacama (Northern Chile). Advances in Bioscience and Biotechnology, 4, 830-839. doi: 10.4236/abb.2013.48110.

1. INTRODUCTION

There is now a worldwide crisis due to the need of new sources of energy. Various disciplines are looking for new conventional and renewable materials to generate heat energy [1]. Among these, the use of microalgae biomass to produce oil and biodiesel is considered increasingly relevant [2-4]. Biodiesel generated from algae offers many advantages, e.g. during algae photosynthesis the atmospheric CO2 is fixed at a similar rate to that produced during combustion process of biodiesel [5,6], thus, the balance of CO2 pollution generated is virtually zero. In addition, microalgae show a high productivity rate with yields per acre nearly 30 times greater than those of corn or soybeans [7]. However, the oil quality for biodiesel production is not always appropriate because it has low triglycerides content, and high concentration of free fatty acids. The latter can interfere with the trans-esterification process required for the biofuel obtaining [4]. Different culture tests including high salinity and nutrient limitation (e.g. Mg, K, Na, NO3 and PO4) can enhance lipid synthesis and particularly triglycerides, but unfortunately these conditions produced a microalgae biomass reduction and therefore, the biodiesel yield also decreases [8-16]. The objective of this study was the assay of specific conditions in salinity and nitrogen deficiency and air bubbling that could stimulate the triglycerides production, while the negative effect over the biomass is mitigated. This will be carried out by a strain of Dunaliella salina SA32007, isolated from the Salar de Atacama, II Region—Chile, in order to potentiate biotechnological properties of the local microorganisms.

Dunaliella salina is a cosmopolitan species living in hypersaline aquatic environments [17]. In Chile there are records of Dunaliella salina in saline lagoons located in Highland Andean Salars, as in Salar de Atacama (Northern, Chile) [18]. High sunlight intensities stimulate this microalgae to produce a variety of carotenoids, being the most abundant in ß-carotene. This pigment is formed as a protective barrier against oxidation due to free radicals caused by sunlight [19]. It is noteworthy that Dunaliella species are recognized primarily for its ability to produce ß-carotene, which is verified by reviewing the vast scientific literature related to this topic [20-23]. Other potentialities of Dunaliella, such as their ability to induce lipids bioaccumulation in large quantities, have also been studied [24].

2. MATERIAL AND METHODS

2.1. Dunaliella salina SA32007 and Media

D. salina SA32007 strain, isolated from the Salar de Atacama (23˚30'0"S, 68˚15'0"W) II Region Antofagasta, Chile, was cultured in two modified versions of Johnson medium [25], one in which the NaCl concentration changed to vary the salinity (identified from here and beyond as “Medium A”) and another that was modified in its N/P ratio (identified from here and beyond as “Medium B”).

“Medium A” contained the following (per liter): NaCl 11.68 g (0.2 M)/29.2 g (0.5 M)/58.4 g (1 M)/116.9 g (2 M)/175.2 g (3 M)/233.6 g (4 M) y 292 g (5 M); MgCl2·H2O 1.5 g; MgSO4·7H2O 0.5 g; KNO3 1 g; KCl 0.2 g; CaCl2 0.2 g; KH2PO4 0.071 g (14/1 N/P Ration, Johnson Medium reference [25]); Tris (hydroxymethyl) aminomethane 2.45 g; EDTA·2Na 1.89 mg; ZnSO4·7H2O 0.087 g; H3BO3 0.61 mg; CoCl·6H2O 0.015 mg, CuSO4·5H2O 0.06 mg; MnCl2 0.23 g, (NH4)6Mo7O24·4H2O 0.38 mg; Fe(III)·EDTA 3.64 mg; vitamin B1 0.1 mg; vitamin B12 0.5 mg; biotin 0.5 mg, pH was fixed to 8.0 with 1 N HCl. Additionally, in this step was tested the 25/1 N/P ration (KH2PO4 0.04 g), corresponding to these determined in the collection strain area.

“Medium B” contained the following (per liter): KH2PO4 2 g/0.2 g/0.071 g/0.04 g and KNO3 1 g (which resulted in a change in the N/P ration equivalent to 0.5/1, 5/1, 14/1, 25/1, respectively); MgCl2·H2O 1.5 g; MgSO4·7H2O 0.5 g; NaCl 29.2 g; KCl 0.2 g; CaCl2 0.2 g; Tris (hydroxymethyl) aminomethane 2.45 g; EDTA·2Na 1.89 mg; ZnSO4·7H2O 0.087 g; H3BO3 0.61 mg; CoCl·6H2O 0.015 mg, CuSO4·5H2O 0.06 mg; MnCl2 0.23 g, (NH4)6Mo7O24·4H2O 0.38 mg; Fe(III)·EDTA 3.64 mg; vitamin B1 0.1 mg; vitamin B12 0.5 mg; biotin 0.5 mg, pH was fixed to 8.0 with 1 N HCl. The N/P rations employed in the “Medium B” were modified from the original recipe of Johnson Medium [25], which presents a N/P ratio of 14/1 corresponding to KH2PO4 0.071 g, and KNO3 1 g.

Each culture medium was supplemented with chloramphenicol 1%, in order to obtain axenic monocultures.

Experiments were conducted under controlled light and temperature conditions at 30˚C with an incident white light of 360 μmol/m2/s and a photoperiod cycle of 15/9 (light/dark). A high light intensity was selected to maintain the optimal conditions for microalgae acclimated to large amounts of light. In fact, measurements of the light incident on the Salar de Atacama often exceed 500 μmol/m2/s [26].

2.2. Experimental I: Variation of Salt Concentration of NaCl

A preliminary screening of salinity-tolerance was performed because the microalgae strain was kept under non-hipersaline culture conditions for more than three years. Fourteen flasks (E-I1-E-I14) with 200 mL of “Medium A” were inoculated with 10 mL of axenic culture of D. salina SA32007 (average cell density of 8.0 × 106 cells/L) and incubated at 30˚C for 30 days. Salinity range varied between 0.2 M and 5.0 M NaCl as described in Dunaliella salina SA32007 and Media section. Control conditions were set according to the isolation place of the microalgae strain, corresponding to a N/P ratio of 25/1 and a salinity level of 0.5 mol/L of NaCl (Table 1).

2.3. Experimental II: Variation of N/P Ratio, Air Bubbling and Nitrogen Deficiency

In this section different N/P ratios were assayed (Table 2) keeping fixed the salinity at 0.5 mol/L NaCl, which yielded the maximum biomass value in the experimental I. Additionally, two other variables were included in this step: the air supply using bubbling and the point within the growth curve of the strain where the nitrogen deficiency was applied. In the second case, two treatments were tested: one in which the deficit of nitrogen was applied Before the Latency Phase (BLP-treatments) and another in which the deficit was applied after the latency Phase (ALP-treatments). As regards the air bubbling, treatments with bubbling (Bb-treatments) and without

Table 1. Experimental I: Provided by varying the salt concentration of NaCl. Experimental II: Provide by varying the N/P ratio, the bubbling and the time of application of nitrogen deficiency.

*Control conditions according to the isolation place of the D. salina SA32007.

Table 2. Experimental II: Provide by varying the N/P ratio, the bubbling and the time of application of nitrogen deficiency.

Abbreviators: Bubbled (Bb); Not Bubbled (NBb); Before Lag Phase (BLP); After Lag Phase (ALP). *Control conditions according to the isolation place of the D. salina SA32007, varying the air bubbling.

bubbling (NBb-treatments) were tested. These treatments were numerated from E-II3 to E-II14 (Table 2). The initial density of each culture averaged 1147 ± 946 Cells/L. The controls (E-II1 and E-II2) were cultured using the same conditions found at the isolation site of the strain, but varying the bubbling (Table 2).

2.4. Biomass Quantification

Microalgal biomass was daily count in duplicate using Neubauer chambers. Microalgae samples were pre-fixed with a 5% Lugol solution. The maximum number of cells in each culture, measured as the carrying capacity (K), was correlated to its intrinsic growth rate (μ).

2.5. Extraction and Quantification of Total Extractable Lipids and Triglycerides

Lipids from the microalgae pellet obtained after biomass centrifugation, were extracted using the Bligh and Dyer method [27]. Thus, the results represent the Percentage of total extractable lipids (%TELs) from fresh biomass of the microalgae. The Percentage of Triglycerides (%TAGs) in total extractable lipids was determined by the glycerol-3-phosphate-oxidase-p-chlorophenol enzymatic method (Triglyceride GOP/PAP AA from Wiener) using triolein as standard [10,28].

Each experimental result was obtained through a triplicated set of data. Standard deviation was calculated in each case. Additionally, the overall effect of different variable tested in this study was statistically validated through the non-parametric significance test of MannWhitney U [29].

3. RESULTS

3.1. Microalgae Growth Curve and Salinity Assays

Figures 1 (a) and (b) show the growth curves of D. salina SA32007 in Media A and B, respectively. Both cultures (N/P ratio of 14/1 and 25/1) showed that microalgae samples cultured with NaCl levels from 0.2 mol/L to 1 mol/L maintained a lag phase for a period between 7 and 11 days, followed by the corresponding exponential phase. For treatments with salinities higher than 1 mol/L of NaCl (E-I3-E-I7 and E-I10-E-I14) no culture growth was appreciated and the mean cell density values of the initial inoculum remained unchanged; this was observed regardless of their N/P ratio. The kinetic behaviour of each treatment is summarized in Table 3. The maximum growth was achieved using 0.5 mol/L of NaCl and a N/P ratio of 14/1 and reached a value of 8.2 × 109 Cells/L (Table 3).

3.2. N/P Ratio, Air Bubbling and Nitrogen Deficiency

Figure 2(a) indicates the growth curves for microalgae cultures subjected to nitrogen deficiency, applied before and after the lag phase with air bubbling. It was observed

Conflicts of Interest

The authors declare no conflicts of interest.

References

[1] Pulz, O. and Gross, W. (2004) Valuable products from biotechnology of microalgae. Applied Microbiology and Biotechnology, 65, 635-648. doi:10.1007/s00253-004-1647-x
[2] Qiang, H., Sommerfeld, M. and Jarvis, E. (2008) Microalgal: Triacylglycerols as feedstocks for biofuel production: Perspective and advances. The Plant Journal, 54, 621-639. doi:10.1111/j.1365-313X.2008.03492.x
[3] Song, D., Fu, J. and Shi, D. (2008) Exploitation of oil-bearing microalgae for biodiesel. Chinese Journal of Biotechnology, 24, 341-348. doi:10.1016/S1872-2075(08)60016-3
[4] Chisti, Y. (2007) Biodiesel from microalgae. Biotechnology Advances, 25, 294-306. doi:10.1016/j.biotechadv.2007.02.001
[5] Kishimoto, M., Okakura T., Nagashima, H., Minowa, T., Yokoyama, S. and Yamaberi, K. (1994) CO2 fixation and oil production using micro-algae. Journal of Fermentation and Bioengineering, 78, 479-481. doi:10.1016/0922-338X(94)90052-3
[6] Brown, L. and Zeiler, K. (1993) Aquatic biomass and carbon dioxide trapping. Energy Conversion and Management, 34, 1005-1013. doi:10.1016/0196-8904(93)90048-F
[7] Metting, F.B. (1996) Biodiversity and application of microalgae. Journal of Industrial Microbiology, 17, 477-489. doi:10.1007/BF01574779
[8] Hui, C. and Jiang, J. (2009) Mini-review. Osmotic responses of Dunaliella to the changes of salinity. Journal of Cellular Physiology, 219, 251-258. doi:10.1002/jcp.21715
[9] Li, Y.Q., Horsman, M., Wang, B. and Wu, N. (2008) Effects of nitrogen sources on cell growth and lipid accumulation of green alga Neochloris oleabundans. Applied Microbiology and Biotechnology, 81, 629-636. doi:10.1007/s00253-008-1681-1
[10] Mendoza, H. and Molina, C. (2008) Variación cuantitativa de la composición en ácidos grasos de Crypthecodinium conil en condiciones de supresión de nitrógeno. Grasas y aceites, 59, 27-32.
[11] Takagi, M., Karseno, Y. and Yoshida, T. (2006) Effect of salt concentration on intercellular accumulation of lipids and triacylglyceride in marine microalgae Dunaliella cells. Journal Bioscience and Bioengineering, 101, 223-226. doi:10.1263/jbb.101.223
[12] Meyer, B. and Oettl, B. (2005) Effects of short-term starvation on composition and metabolism of larval Antarctic krill Euphausia superb. Marine Ecology Progress, 292, 263-270. doi:10.3354/meps292263
[13] Kumar, S., Uma, L. and Subramanian, G. (2003) Nitrogen stress induced changes in the marine cyanobacterium Oscillatoria willei. BDU 130511. FEMS Microbiology Ecology, 45, 263-272. doi:10.1016/S0168-6496(03)00162-4
[14] Yamaberi, K., Takagi, M. and Yoshida, T. (2000) Limited feeding of potassium nitrate for intracellular lipid and triglyceride accumulation of Nannochloris sp. UTEX LB1999. Applied Microbiology and Biotechnology, 54, 112-117. doi:10.1007/s002530000333
[15] Xing-Qing, X. and Beardall, J. (1997) Effect of salinity of fatty acid composition of a green microalga from an Antarctic hypersaline lake. Photochemistry, 45, 655-658. doi:10.1016/S0031-9422(96)00868-0
[16] Garnham, G.W., Codd, G.A. and Gadd, G.M. (1992) Kinetic of uptake and intracellular location of cobalt, manganese and zinc in the estuarine green alga Chlorella salina. Biotechnology, 37, 270-276.
[17] Williams W.D. (1998) Salinity as a determinant of the structure of biological communities in salt lakes. Hydrobiology, 381, 191-201. doi:10.1023/A:1003287826503
[18] González, J. and Pena, A. (2002) Estrategia de adaptación de microorganismos halófilos y Debaryomyces hansenii (Levadura halófila). Revista Latinoamericana de Microbiología, 44, 44-45.
[19] Harwood, J. and Guschina, I. (2009) The versatility of algae and their lipid metabolism. Biochemistry, 91, 679-684. doi:10.1016/j.biochi.2008.11.004
[20] Gomez, P., González, M. and Becerra, J. (1999) Quantity and quality of b-carotene produces by two strains of Dunaliella salina (TEODORESCO 1905) from the Northern Chile. Boletín de la Sociedad Chilena de Química, 44, 463-468.
[21] Kopecky, J., Schoefs, B., Loest, K., Stys, D. and Pulz, O. (2000) Microalgae as a source for secondary carotenoid production: A screening study. Algological Studies, International Journal of Phycological Research, 133, 153-168.
[22] Orosa, M., Torres, E., Fidalgo, P. and Abalde, J. (2000) Production and analysis of secondary carotenoids in green algae. Journal of Applied Phycology, 12, 553-556. doi:10.1023/A:1008173807143
[23] Ben-Amotz A. (2004), Industrial production of microalgal cell-mass and secondary products-major industrial species. Handbook of Microalgal Culture: Biotechnology and Applied Phycology, 273.
[24] González de Molina, Dr. M. (2011) Technical paper SEAE: Series introduction to agroecology: The fertilization and balance of nutrients in agroecological systems. Spanish Society for Organic Farming (SEAE), Catarroja, Valencia, 1-24.
[25] Johnson, M.K., Johnsone, J., Macelroyr, L., Speerh, L. and Bruffb, S. (1968) Effects of salts on the halophilic alga Dunaliella viridis. Journal of Bacteriology, 95, 1461-1468.
[26] Kampf, S.K. and Tyler, S.W. (2005) Spatial characterization of evaporation and land surface energy fluxes at the Salar de Atacama, Northern Chile using ASTER image classification. Advances in Water Resources, 29, 336-354. doi:10.1016/j.advwatres.2005.02.017
[27] Bligh, E.G. and Dyer, W.J. (1959) A rapid method for total lipid extraction and purification. Canadian Journal Biochemistry and Physiology, 37, 911-917. doi:10.1139/o59-099
[28] Ben-Amontz, A., Tornabene, T. and Thomas, W. (1985) Chemical profile of selected species of microalgae with emphasis on lipids. Journal Phycology, 21, 72-81. doi:10.1111/j.0022-3646.1985.00072.x
[29] Mann, H.B. and Whitney, D.R. (1947). On a test of whether one of two random variables is stochastically larger than the other. Annals of Mathematical Statistics, 18, 50-60. doi:10.1214/aoms/1177730491
[30] Serpa, R. and Calderon, A. (2005) The effect of salinity stress in forth strains of Dunaliella salina TEOD in Perú. Ecología Aplicada, 4, 127-133.
[31] Joyce, L., Stuart, D. and Duncan, A. (1977) The salt relations of Dunaliella. Microbiology, 113, 131-138.
[32] Cifuentes, A., González, M., Conejeros, M., Dellarosa, V. and Parra, O. (1992) Growth and carotenogenesis in eighth straits of Dunaliella salina teodoresco from Chile. Journal of Applied Phycology, 4, 111-118. doi:10.1007/BF02442459
[33] Ben-Amontz, A. and Avron, M. (1975) Adaption of the unicellular alga Dunaliella parva to a saline environment. Journal of Phycology, 11, 50-54.
[34] Oren, A. (2005). A hundred years of Dunaliella research: 1905-2005. Saline Systems, 1, 1-14. http://www.aquaticbiosystems.org/content/1/1/2
[35] Alonso, H. and Risacher, F. (1996) The Salar de Atacama Geochemistry, part 1: The origen of components and sa-line balance. Revista Geológica de Chile, 23, 113-122.
[36] Hammes, F., Vital, M. and Egli, T. (2010) Critical evaluation of the volumetric “bottle effect” on microbial batch growth. Applied Environmental Microbiology, 76, 1278-1281. doi:10.1128/AEM.01914-09
[37] Lara, M. and García, E. (2007) La utilidad experimental y los procesos desencadenados por el sometimiento de los organismos a condiciones de hambruna. Encuentros en la Biología, 121, 2. http://www.encuentros.uma.es/encuentros121/121.pdf
[38] Xin, L., Hong-Ying, H. and Yu-Ping, Y. (2011) Growth and lipid accumulation properties of a freshwater microalga, Scenedesmus sp. under different cultivation temperature. Bioresource Technology, 102, 3098-3102. doi:10.1016/j.biortech.2010.10.055
[39] Cohen, Z. and Khozin-Golberg, I. (2000) The role of triacylglycerol as a reservoir of polyunsaturated fatty acids for the rapid production of chloroplastic lipids in certain microalgae. Biochemical Society Transactions, 28, 740-743. doi:10.1042/BST0280740
[40] Mendoza, H., Martel, A., Jiménez del Río, M. and García, G. (1999) Oleic acid is the main fatty acid related with carotenogenesis in Dunaliella salina. Journal of Applied Phycology, 11, 15-19. doi:10.1023/A:1008014332067
[41] Hernández, L., Quintana, M. and Morris, H. (2000) Obtención de glicerol a partir de la microalga Dunaliella salina. Revista Cubana de Farmacia, 34, 134-137.
[42] Ben-amontz, A., Katz, A. and Avron, M. (1982) Accumulation of beta-carotene in halotolerant algae: Purification and characterization of beta-carotene rich globules from Dunaliella bardawill (Chlorophyceae). Journal of Phycology, 18, 529-537. doi:10.1111/j.1529-8817.1982.tb03219.x
[43] Zhila, N., Kalacheva, G. and Volova, T. (2005) Influence of nitrogen deficiency on biochemical composition of the green alga Botryococcus. Journal of Applied Phycology, 17, 309-315. doi:10.1007/s10811-005-7212-x
[44] Yamaberi, K., Takagi, M. and Yoshida, T. (1998) Nitrogen depletion for intracellular triglyceride accumulation to enhance liquefaction yield of marine microalgal cell into fuel oil. Journal of Marine Biotechnology, 6, 44-48.
[45] Fidalgo, J.P., Cid, A., Torres, E., Sukenik, A. and Herrero, C. (1998) Effects of nitrogen source and growth phase on proximate biochemical composition, lipid classes and fatty acid profile of the marine microalga Isochrysis galbana. Aquaculture, 166, 105-116. doi:10.1016/S0044-8486(98)00278-6
[46] Giordano, M. and Beardall, J. (2009) Impact of environmental conditions on photosynthesis, growth and carbon allocation strategies of hypersaline species of Dunaliella. Global NEST Journal, 11, 79-85. http://www.gnest.org/journal/Vol11_No1/79-85_566_Giordano.pdf

Copyright © 2024 by authors and Scientific Research Publishing Inc.

Creative Commons License

This work and the related PDF file are licensed under a Creative Commons Attribution 4.0 International License.