Вы находитесь на странице: 1из 9

Algal Research 12 (2015) 368–376

Contents lists available at ScienceDirect

Algal Research

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

Isolation of prospective microalgal strains with high saturated fatty acid


content for biofuel production
A.V. Piligaev a,⁎, K.N. Sorokina a,b,c, A.V. Bryanskaya b, S.E. Peltek b, N.A. Kolchanov b,c, V.N. Parmon a,c
a
Boreskov Institute of Catalysis, pr. Lavrentieva 5, Novosibirsk 630090, Russia
b
Federal Research Center Institute of Cytology and Genetics of the Siberian Branch of the RAS, pr. Lavrentieva 10, Novosibirsk 630090, Russia
c
Novosibirsk State University, Pirogova str. 2, Novosibirsk 630090, Russia

a r t i c l e i n f o a b s t r a c t

Article history: Isolation of new microalgal strains with unique and valuable properties from diverse environmental conditions is
Received 3 March 2015 an important starting point for the establishment of high quality feedstock for biofuel production. In this study,
Received in revised form 20 July 2015 we have isolated twelve strains of microalgae, belonging to the genera Chlorella, Botryococcus, Scenedesmus,
Accepted 31 August 2015
Nannochloris and Bracteacoccus from samples obtained from various natural sources by flow cytometry. Five of
Available online xxxx
them were selected for further evaluation, as lipid producers on the basis of Nile Red fluorescent microscopy
Keywords:
screening. Analysis of the biomass in the exponential and stationary phases of growth showed that the strain
Microalgae Scenedesmus abundans A1175 has the highest specific growth rate (0.20 ± 0.01 d−1), biomass productivity
Biofuel (73.82 ± 4.53 mg L−1 d−1) and lipid content (44.4 ± 2.7%) in nitrogen depleted conditions. Strains Chlorella
Isolation of strains vulgaris A1123 and S. abundans A1175 have a high total content of saturated (SFA) and monounsaturated
Identification (MFA) fatty acids (67.0% and 72.8% respectively). S. abundans A1175 also had low polyunsaturated fatty acid
Fatty acid (PUFA) content that would allow for its use as a source of high quality biofuels.
Lipid © 2015 Elsevier B.V. All rights reserved.

1. Introduction standards EN 14214:2003 and ASTM D6751 is a low content of polyun-


saturated fatty acids in raw oils used a feedstock. Biofuels that contain
Microalgae are widespread in nature, including soil, freshwater and components with more saturated bonds are more resistant to oxidation,
marine environments [1]. Microalgae biomass has been considered as a and at combustion, glycerides remaining in the fuel are not polymer-
promising source for production of motor fuels in comparison with ized, which increases the reliability of diesel engines [7]. Thus, to find
other sources of renewable biomass [2]. It was shown that oil produc- a prospective microalgae strain for the production of high-quality biofu-
tivity of microalgae is 100 times higher than that of some agricultural el, it is necessary to screen for natural habitat strains with a high content
crops, such as canola, soy and oil palm [3,4]. Currently, microalgal of SFA and MFA with no additional requirements for expensive supple-
biomass is used to produce third generation biofuels, biogas and ments for fast growth (e.g. vitamins). A number of studies have shown
bioethanol, but is also considered as a source of other valuable compo- that the total content of SFA and MFA can reach high values (60–68% in
nents: protein, polyunsaturated fatty acids, pigments, sugars and anti- some strains like in Chlorella vulgaris ESP-31 [8]). However, features of
biotics [5]. such strains remain poorly understood in terms of the physiology of
High oil content of some microalgal strains makes them suitable for the accumulation and composition of lipids, and their applicability to
biodiesel production. Microalgae belonging to such genera as Chlorella, the production of valuable by-products, mainly protein and
Dunaliella, Nannochloris, Nannochloropsis, Neochloris, Porphyridium, and carbohydrates.
Scenedesmus contain 20–50% of lipids by weight, making them suitable In this study, we have isolated a number of microalgal strains from
for biodiesel production [4]. Nevertheless, it was shown that strains various natural sources. The biomass accumulation, productivity, lipid
with high lipid content have a low biomass productivity that significant- content and composition of the isolated strains were studied during
ly reduces their effectiveness as lipid producers [6]. In addition to high the exponential and stationary phases of autotrophic growth. Using
lipid content, prospective microalgal strains need to have high growth flow cytometry, a number of microalgal strains were isolated and iden-
rates and a specific composition of produced lipids. An important cri- tified. Five of them were selected through screening with Nile Red mi-
terion for production of high quality biodiesel that meets biodiesel croscopy as lipid producers. Analysis of their productivity for biomass
and lipids, lipid accumulation and biochemical composition revealed
that two strains, S. abundans A1175 and C. vulgaris A-1123, have optimal
⁎ Corresponding author. characteristics for production of high quality oil with high total content
E-mail address: piligaev@catalysis.ru (A.V. Piligaev). of MSA and SFA.

http://dx.doi.org/10.1016/j.algal.2015.08.026
2211-9264/© 2015 Elsevier B.V. All rights reserved.
A.V. Piligaev et al. / Algal Research 12 (2015) 368–376 369

2. Materials and methods 0.08, and cultured at 150 rpm in an orbital shaker at 28 °C as described
until the stationary phase of growth was reached.
2.1. Isolation of microalgae by cell sorting with flow cytometry
2.5. Determination of dry cell weight
Samples of soil and water were inoculated in 50 cm3 Bold's Basal Me-
dium (BBM) [9] and cultivated for two weeks at a temperature of 28 °C, The relationship between OD and dry cell weight (DW) was deter-
under a 16 h-light/8 h-dark cycle, and illumination of 60 μmol m−2 s−1. mined for each strain. During the exponential phase of growth, the OD
Microalgae cells were separated from the contaminating microorgan- was measured at 680 nm, and 10 cm3 of liquid was withdrawn from cul-
isms by flow cytometry. Cell sorting was performed on BD FACSAria ture (not less than 5 times at different time points in triplicate) and
flow cytometer, and a sapphire laser with a wavelength of 488 nm transferred on a glass fiber filter, dried at 105 ºС overnight and weighed.
was used for excitation of chlorophyll fluorescence. The fluorescence in- The relationship between OD and DW was determined through linear
tensity was determined by photodetector with a dichroic mirror 655LP regression and converted to g L−1 with R2 N 0.98.
and a narrowband optical filter 695/40 nm. FACSFlow™ was used as a
compressing fluid in the sorting process. Before sorting, suspension
2.6. Evaluation of growth rates and productivity parameters
was purified from large aggregates, using cell strainers (35 μm, BD Fal-
con). Microalgae cell fraction had at least 6 × 104 cells that were plated
Biomass productivity (BP) was calculated through the equation BP
onto BBM agar and cultured as described until visible colonies were ob-
(mg L−1 d−1) = (X2 − X1) · (t2 − t1)−1 where X1 and X2 were the bio-
served. The purity of microalgal cultures was verified by microscopy
mass dry weight concentrations (mg L−1) on days t1 (start of the expo-
and by the absence of bacterial and fungal growth on LB and PDA agar
nential phase) and t2 (end of the exponential phase), respectively.
layered with diluted microalgae cultures from BBM agar plates.
Specific biomass growth rate (μ) was calculated in the exponential
growth phase by μ (d−1) = (lnX2 − lnX1) · (t2 − t1)−1 where X1 and
2.2. Identification of the isolated strains of microalgae and X2 are the lipid dry weight concentrations (mg L−1) on days t1 (start
phylogenetic analysis of exponential growth) and t2 (end of exponential growth),
respectively.
Genomic DNA from microalgal cultures was isolated using the Wiz- Doubling time (DT) is defined as DT = ln2 · (μ) −1.
ard SV Genomic DNA Purification System (Promega, USA) according to All experiments were carried out in triplicate, and data are expressed
the manufacturer's instructions. The partial sequence of the 18S rRNA as mean ± SD.
gene was amplified from the 50 ng genomic DNA by PCR by Taq poly-
merase using universal primers 5′-ACCTGGTTGATCCTGCCAGT-3′ 2.7. Fluorescent microscopy of cells after Nile Red staining
(forward) and 5′-TCAGCCTTGCGACCATAC-3′ (reverse) [10], PCR was
performed under the following conditions: 1 cycle 95 °C — 3 min, 40 cy- Fluorescent staining of 1 cm3 of microalgal cell suspension (105–106
cles (95 °C — 30 s, 54 °C — 20 s, and 72 °C —1 min 30 s), 1 cycle — 72 °C cells) was performed using 0.05 cm3 of the Nile Red solution in acetone
for 10 min and products were analyzed by electrophoresis. DNA se- (0.1 mg cm−3). The mixture was incubated for 10 min at 37 °C in dark-
quencing was performed with BigDye® Terminator v3.1 Kit (Applied ness. The fluorescence of the cells of stained and control samples of
Biosystems) and the nucleotide sequence was determined on ABI microalgae were observed on Axioskop 2 Plus (Carl Zeiss) in the Inter-
3730XL sequencer (Applied Biosystems). The obtained sequences institutional Shared Center for Microscopic Analysis of Biological
were aligned with BLAST to sequences of microalgae 18 s rRNA from Objects SB RAS. Zeiss filter sets FS10 (excitation bandpass, 450 to
GenBank database [11]. Multiple sequence alignment and phylogenetic 490 nm; emission bandpass, 515 to 565 nm) and FS20 (excitation
tree construction were performed using MEGA 5 [12]. The phylogenetic bandpass, 546/12 nm; emission bandpass, 575 to 640 nm) were used
tree was built using the Neighbor-Joining algorithm [13], and the statis- for fluorescence detection. Images were taken with Axiovision imaging
tical reliability of its topology was determined by the Bootstrap test [14]. v4.6. Acquisition and processing of data was done using the Axiovision
v4.6 software.
2.3. Preliminary screening for lipid production in microalgae
2.8. Determination of Nile Red fluorescence
Equal amounts of cells from each microalgae isolates obtained dur-
ing the isolation were cultured in 24 well plates in 1.5 cm3 BBM-3 N me- Fluorescent staining of neutral lipids of microalgae was performed
dium supplemented with 0.75 g L−1 NaNO3 under a 16 h-light/8 h-dark using the Nile Red dye. For this purpose, 0.025 cm3 of Nile Red in ace-
cycle illumination at 60 μmol m−2 s−1 in 2% CO2 atmosphere for 7 days tone (0.1 mg cm−3) was added to 0.5 cm3 of the cell suspension con-
and were agitated in an orbital shaker at 150 rpm. After incubation, cells taining 20% DMSO, with the concentration of cells normalized at OD
were harvested by centrifugation at 1000 g for 5 min, transferred to the 0.2 (680 nm). The mixture was incubated for 10 min at 37 °C in dark-
BBM medium without NaNO3, and similarly cultured in 24 well plates ness. Analysis was performed with Perkin-Elmer EnVision2103
for up to 10 days to trigger lipid accumulation in cells. During cultivation Multilabel Reader using 525/579 excitation/emission filters. The relative
in the BBM medium, the pattern of accumulation of lipids was observed fluorescence intensities were obtained by subtracting of both the auto-
by fluorescent microscopy with Nile Red cell staining. fluorescence of unstained algal cells and fluorescence intensity of Nile
Red of blank medium containing DMSO and Nile Red Dye from the fluo-
2.4. Cultivation of microalgae rescence intensities measured in the stained cells.

A single colony of each strain cultivated on BBM agar was picked up 2.9. Analysis of nitrate concentration in medium
and inoculated into 50 cm3 Erlenmeyer flacks in 25 cm3 BBM media and
cultivated at 80 rpm agitation on an orbital shaker at 28 °C (without CO2 Nitrate concentration during the cultivation of microalgal strains in
supplementation), at an illumination of 60 μmol m−2 s−1 under a 16 h- medium was measured as described in [15]. Samples of cultural liquid
light/8 h-dark cycle until the mid-exponential phase was reached. Cul- were centrifuged at 3000 g for 5 min. The supernatant was collected
ture growth was monitored by measuring the optical density (OD) at and the absorbance was measured at 220 nm. Sodium nitrate solution
680 nm using a UV–VIS spectrophotometer (Epoch, Biotek). Cells were with concentrations of up to 500 μM were used as a standard in all
inoculated in 250 cm3 conical flasks, containing 100 cm3 of BBM at OD measurements.
370 A.V. Piligaev et al. / Algal Research 12 (2015) 368–376

2.10. Analysis of total lipid content and productivity 2.14. Determination of carbohydrate content

For lipid analysis, cells were pelleted by centrifugation at 3000 g for The carbohydrate content of the microalgal biomass was deter-
10 min, washed twice with milli-Q water and then lyophilized in a mined with the phenol-sulfuric acid method [17]. For this purpose,
freeze drier (FreeZone, Labconco) and weighed. Dry biomass was used 10 mg of lyophilized algae was added to 10 cm 3 of water. After
for experiments and all measurements were done in triplicate. For this, a 1 cm3 aliquot of sample was added to 3 cm3 of sulfuric acid
lipid extraction up to 100 mg lyophilized algae biomass was dissolved and 1 cm 3 of 5% aqueous solution of phenol, and the mixture was
in 2 cm3 of chloroform:methanol (2:1), and sonicated for 2 min at stirred and incubated for 5 min at 90 °C. The amount of carbo-
130 W with CPX 130 Ultrasonic Processor (Cole Parmer, USA). The mix- hydrates was determined using a wavelength of 488 nm against a
ture was further added with 0.25 volume of 0.9% NaCl solution and calibration curve based off of a known concentration of glucose. Ex-
stirred vigorously. After phase layering, the organic phase was evaporat- periments were carried out in triplicate, and the data are expressed
ed in a stream of nitrogen, and the sample was weighed. Lipid content as mean ± SD.
was determined as a ratio of weight of lipid extract to weight of dry bio-
mass. Lipid extracts were also used for the analysis of fatty acids and
lipid composition. 3. Results and discussion
Lipid productivity was calculated as (mg L−1 d− 1) = (X2 −
X1) · (t2 − t1)−1 where X1 and X2 were the lipid dry weight concentra- 3.1. Isolation of microalgae with flow cytometry
tions (mg L−1) for days t1 (start of the exponential growth phase) and t2
(end of the exponential growth phase), respectively. Experiments were In this study, more than 100 environmental samples were cultivated
performed in triplicate, data are shown as mean ± SD. from different sources (the salt lakes of Novosibirsk and Altai regions,
the anthropogenic ecosystems of Kemerovo oblast (Western Siberia),
2.11. TLC of microalgal lipid extract and the terrestrial hot springs in the Uzon Caldera and Valley of Geysers
(Kamchatka) for the isolation and screening of microalgae strains. Only
To analyze the lipid composition in microalgal biomass, thin layer 49 of the samples revealed growth of microalgaе and were used in our
chromatography (TLC) was performed on TLC silica gel 60G plates work to obtain microalgae cultures.
(Merck). 0.003 cm3 of extracted lipid solution (10 mg in 1 cm3 of chlo- In previous studies, it was shown [18–20] that flow cytometry can be
roform) was layered on 10 × 15 cm plates, air-dryed and separated successfully applied to reduce contamination of the isolated microalgae
using hexane-diethyl ether-acetic acid (80:20:1) as an eluent mixture. cultures by extraneous microflora through the separation of subpopula-
A mixture (10 mg cm−3) of lipids, including a mono-, di-, and triglycer- tions of cells in the suspension. In our study, for the detection of
ide mix (Sigma) and oleic acid (Sigma) was used as a standard. Spots microalgae cells during sorting, we measured the following parameters:
were detected after spraying the plate with 1% solution of phospho- the red fluorescence of chlorophyll а (N650 nm) and forward light scat-
molybdic acid in ethanol and heating at 120 °C for 15 min. tering, which indicates cell size. Fig. 1 shows typical cytograms of the
microalgae culture obtained from natural samples before and after
sorting.
2.12. GC–MS analysis of FAME During cell sorting, we observed two or more different cell
clusters on the cytogram. Clusters with the highest intensity of fluo-
To analyze the composition of fatty acids in the microalgal oil, the rescence and the highest forward scattering corresponed to
lipids were esterified with methanol to obtain fatty acid methyl esters microalgae cells, while clusters with low intensity of fluorescence
(FAME) [16]. To avoid the oxidation of unsaturated fatty acids, all sam- and low scattering probably contained non-photosynthetic cells
ples were processed under a nitrogen atmosphere. For this reaction, up that may be normally present in the natural samples. Red color (re-
to 2 mg of lipids was added to 0.3 cm3 of a 2% H2SO4 in methanol solu- gion P1) in the Fig. 1 (a and b) shows the fraction with the highest
tion and incubated for 1.5 h at 80 °C. 50 μg of heptadecanoic acid C17:0 forward scattering and absorption in the area of 695 nm, which is
(Sigma) was used as an internal standard for esterification and added to characteristic of microalgae cells. After flow sorting, 12 unique
the mixture prior to reaction. 0.3 cm3 of 0.9% (w/v) NaCl and 0.3 cm3 of microalgal strains were isolated and maintained as the axenic
hexane were then added, vortexed and separated by centrifugation at cultures.
3000 g for 5 min. The upper hexane layer was used for GC–MS analysis.
The composition of fatty acid methyl esters was analyzed by GC–MS
on Agilent 7000B with electron impact ionization (70 eV) on a ZB-WAX 3.2. Taxonomic identification of microalgae
column (30 m × 0.25 mm ∗ 0.25 μm), with a temperature gradient from
100 °C to 260 °C at an increase rate of 12 °C min−1. Injector temperature According to 18S rRNA analysis, the isolated microalgal strains be-
was set at 260 °C. Carrier gas (helium) flow was set to 1.2 cm3 min−1. long to the genera Chlorella, Botryococcus, Scenedesmus, Nannochloris
Peaks of fatty acids were identified by using the NIST'11 library, and and Bracteacoccus with more than 99% of homology (Fig. 2).
the relative amount of individual fatty acid was calculated by the inte- It was found that the isolated strains A1123, A1176, A1182 and
grated area percentage from the total amount of fatty acids. Experi- A1135 formed a single closely related cluster and had 99% homology
ments were carried out in triplicate, and the data are expressed as with the strains of C. vulgaris FR865683.1, Chlorella sp. AB713411.1,
mean ± SD. Chlorella sp. AY195981.1 and Chlorella sorokiniana AB731602.1, respec-
tively. Strains A1139 and A1144 had 99% homology with the strains
2.13. Determination of protein content Nannochloris bacillaris AB080300.1 and Bracteacoccus sp. JQ259951.1,
respectively. Strains A1175 and A1167 showed 99% and 100% degree
For protein extraction, 10 mg of lyophilized biomass was added to of similarity of genetic sequences with the strains of S. abundans
10 cm3 of 0.5 M solution of NaOH and incubated for 10 min at 80 °C, X73995.1 and Scenedesmus obliquus FR865738.1 respectively. Strains
and then centrifuged for 5 min at 14,000 g. Protein content in the ob- A1115, A1138, A1113 and A1162 form a single cluster, belonging to
tained extract was examined with the Quick Start kit (Bio-Rad) accord- the genus Botryococcus. Strains A1115 and A1138 had 100% and 99%
ing to the manufacturer instructions and using BSA as a standard. homology with the strain Botryococcus sp. GU250969.1, and strains
Experiments were performed in triplicate, and the data are expressed A1113 and A1162 had 99% homology with the strain of Botryococcus
as mean ± SD. sp. AJ581914.1.
A.V. Piligaev et al. / Algal Research 12 (2015) 368–376 371

Fig. 1. Cytograms of a highly contaminated sample of microalgal cuture before (a) and after (b) cell sorting. The X axis (PerCP-A) corresponds to the fluorescence channel 695/40 nm, and
axis Y (FSC-H) — with forward light scattering. The region P1 corresponds to the signal from microalgae fraction.

3.3. Screening of microalgal stains with high lipid content 3.4. Cultivation of the selected microalgal strains

To screen for prospective microalgal strains with high lipid content, The most important parameters in the screening of microalgal
the accumulation of lipids in the microalgae cells under nitrogen-de- strains for biodiesel production are high lipid content and biomass pro-
pleted conditions for up to 10 days was observed using staining with ductivity [21]. Selected strains were cultured up to 26 days on BBM, and
Nile Red. All of the isolated strains under study produced lipids after in- the following parameters were evaluated: accumulation of lipid drop-
duced starvation, but had different patterns of accumulation of lipid lets in cells stained with Nile Red and analyzed with fluorescent micros-
droplets. For example, the majority of the cells in the cultures of copy (Fig. 3), dry weight of the biomass (Fig. 4), nitrogen concentration
Scenedesmus obliquus A-1167, Botryococcus sp. A-1162 and S. abundans in the medium (Fig. 5) and the dynamics of lipid accumulation shown
A-1175 had synchronic manners of accumulation of lipid droplets. by Nile Red fluorescence (Fig. 6). Growth rates and productivity param-
Other strains visually had fewer lipid-producing cells (approximately eters of the selected strains are presented in Table 1.
25%), except for the strains C. vulgaris A-1123 and C. sorokiniana A- As shown in Fig. 4, the stationary phase of growth is achieved by
1135. More than 50% of the cells in these cultures contained visible strain S. abundans A1175 on 19th day, and this timepoint correlates
droplets. Based on this observation, we have selected strains with the total depletion of nitrogen (Fig. 5). This strain differs from
S. abundans A-1175, S. obliquus A-1167, Botryococcus sp. A-1162, the others because of its coordinated accumulation of biomass and
C. sorokiniana A-1135 and C. vulgaris A-1123 for further analysis. lipids, as was discovered through the growth dynamics of biomass
and Nile Red fluorescence intensity (Figs. 4 and 6). In addition, this
strain has the highest specific growth rate of 0.20 ± 0.01 d−1 and bio-
mass productivity 73.82 ± 4.53 mg L−1 d−1, and, thus, the shortest dou-
bling time of 3.6 ± 0.5. Fluorescent microscopy analysis revealed the
formation of large oil droplets in the stationary phase of growth for
this strain (Fig. 3, BF2) on the 19th day. Also, in stationary phase of
growth, this strain had a higher lipid content (44.4 ± 2.7%) than
Botryococcus sp. A1162 (33.4 ± 2.3%) on 20th day, C. vulgaris A1123
(38.3 ± 1.0%) on 26th day, and this strain also had a lipid productivity
of 32.8 ± 1.5 mg L− 1d− 1. It exceeds the similar values from other
strains significantly; specifically, lipid production in S. abundans A1175
is 2.4 times greater than that of C. vulgaris A1123 and 3.3 times greater
than that of Botryococcus sp. A1162. Thus, S. abundans A1175 has the
greatest ability to produce biomass and lipids in the stationary phase
of growth.
Strains C. vulgaris A1123 and S. obliquus A1167 exhibited a relatively
high lipid content (38.3 ± 1.0% and 41.2 ± 2.2% respectively), that was
also confirmed with fluorescent microscopy (Fig. 3, BF1 and BF4). Nev-
ertheless, compared to S. abundans A1175, both strains have delayed
lipid accumulation as compared to the growth of biomass in the expo-
nential phase (Fig. 6), C. sorokiniana A1135, among the other strains
under study, has the lowest biomass (10.60 ± 0.34 mg L−1 d−1) and
lipid productivity (3.3 ± 0.1 mg L−1 d−1), accompanied by a low lipid
content (31.2 ± 3.4%) and rare oil droplet formation in cells (Fig. 6).
Low lipid productivity for this strain is also correlated with high residual
Fig. 2. Phylogenetic analysis of the isolated strains of microalgae. Numbers represent the nitrogen content in the medium (0.08 g L−1) in the stationary phase of
results of 500 bootstrap replicates. growth (Fig. 5).
372 A.V. Piligaev et al. / Algal Research 12 (2015) 368–376

Fig. 3. Transmitted light and fluorescent images of microalgae stained with Nile Red in exponential and stationary phase of growth. Scale bars show relative size of cells. F — fluorescent
image; T — transmitted light; A — exponential phase; B — stationary phase. 1 — Chlorella vulgaris A1123 (26th day); 2 — Scenedesmus abundans A1175 (19th day); 3 — Chlorella sorokiniana
A1135 (25th day); 4 — Scenedesmus obliquus A1167 (25th day); 5 — Botryococcus sp. A1162 (20th day).

3.5. Analysis of the composition of lipids by TLC 3.6. Analysis of fatty acid composition of the biomass of microalgae at
different stages of growth
The composition lipid extracts from microalgae was studied in the
exponential and stationary growth phases through TLC (Fig. 7). An important parameter that determines the quality of biodiesel is
The stationary phase was indicated by the presence of triglycerides its fatty acid composition. The structure of the fatty acids affects key
that were mostly absent in the exponential phase in all strains except properties of biodiesel, especially its cetane number, flash point, viscos-
for C. sorokiniana A1135, which has the lowest amount of lipids station- ity, oxidation stability, and fluidity at low temperature [3,22]. In this re-
ary phase (31.2 ± 3.4%) and the fewest lipid droplets in cells (Fig. 3, gard, a raw material for biofuel production has to meet a number of
BF3). It can also be noted that C. vulgaris A1123 and S. abundans requirements. For example, quality oil shall contain high levels of satu-
A1175 showed apparent increases in triglyceride content in the station- rated fatty acids and low levels of polyunsaturated ones (with the num-
ary phase compared with the other strains. ber of double bonds being greater than four) [23]. Polyunsaturated fatty
A.V. Piligaev et al. / Algal Research 12 (2015) 368–376 373

Fig. 4. DW of microalgal strains during cultivation.

acids are susceptible to polymerization, and, as such, resins can form tendency to change fatty acid composition during cultivation
through oxidation during storage or through oxidative and thermal po- (Table 2). For example, in the exponential and stationary phases of
lymerization at higher temperatures and pressures during combustion. growth. High levels of palmitic С16:0, stearic С18:0 and oleic С18:1
In addition, biofuels must have a sufficiently high level of saturated fatty acids were observed. However, in the stationary phase, a significant de-
acids to increase the cetane number and, at the same time, a sufficient crease in the content of stearic С18:0 acid was observed in C. vulgaris
level of monounsaturated fatty acids. It is shown that palmitoleic, as A1123 (27.8 ± 4.9%), S. obliquus A1167 (19.7 ± 0.7%) and Botryococcus
well as oleic acid, significantly enhances cold flow properties of biofuels sp. A1162 (20.2 ± 1.5%).
[24]. The amount of oleic acid С18:1 remained stable in all strains in both
Compared to oils obtained from soybeans and canola, oil produced growth phases, except the strains C. vulgaris A1123 and Botryococcus sp.
by microalgae has a higher content of fully saturated fatty acids: A1162 showed a nearly twofold increase in C18:1 levels during the sta-
palmitic C16:0 and stearic C18:0. Soyabean oil and rapeseed oil are char- tionary phase. The amount of n-2 unsaturated fatty acid biomass was
acterized by a high content of linoleic С18:2 and oleic acid С18:1 [25]. significantly different among different strains: in exponential phase,
For the accurate determination of unsaturated fatty acids in our the linoleic acid С18:2 content of C. vulgaris A1123 was 9.5 ± 3.1%,
microalgae cultures, we used GC–MS. All manipulations with lipid ex- S. obliquus A1167, 3.9 ± 0.6% and Botryococcus sp. A1162, 13.1 ± 0.2%.
tracts were conducted under a nitrogen atmosphere. All strains had a During the stationary phase, these values increased by almost two

Fig. 5. Nitrogen concentration during cultivation of microalgae.


374 A.V. Piligaev et al. / Algal Research 12 (2015) 368–376

Fig. 6. Nile Red fluorescence intensity of stained microalgae cells.

times, up to 21.5 ± 3.4%, 6.9 ± 0.8 and 20.2 ± 4.7%, respectively. The compared with all other strains, showed lower levels of n − 2 and
highest content of n-3 PUFA, such as hexadecatrienoic acid С16:3, was n−3 PUFA.
in C. sorokiniana A1135, which had levels of 7.8 ± 0.9% in the exponen- These results are consistent with other studies. For example, C.
tial phase and 6.8 ± 0.3% in the stationary phase. Linolenic acid С18:3 vulgaris A1123 also had a higher content of saturated fatty acids than
content was the highest in the exponential and stationary phases of C. vulgaris YSL04 [26] and Chlorella sp. BR2 [27]. C. vulgaris A1123 also
C. sorokiniana A1135 (14.0 ± 4.0% and 13.8 ± 0.8% respectively) and had comparable biomass productivity to the values for strain C. vulgaris
S. obliquus A1167 (13.4 ± 3.2% and 20.5 ± 2.9 respectively). In addition, 259 that were published in the work [28]. S. abundans A1175, isolated in
high hexadecatetraenoic acid С16:4 levels were found in S. abundans this work, has a little bit smaller lipid productivity and content than the
A1175 (3.6 ± 0.7%) and S. obliquus A1167 (7.5 ± 1.7%) in the exponen- strain S. abundans from [29]. This can be partly explained by its culti-
tial phases, which significantly increased during nitrogen depletion only vation in another medium. Results on the total content of MFA and
for S. obliquus A1167 (to 10.5 ± 0.6%). The latter strain also had a rela- SFA obtained for S. abundans A1175 in this study are comparable to
tively high content of octadecatetraenoic acid С18:4 (2.5 ± 0.4%) during the data from [29], which also investigated the composition of
nitrogen depletion. Thus, the total content of PUFA was maximal in the fatty acids of S. abundans, which showed comparable results for
strain C. sorokiniana A1135 (33.4%) in the stationary phase of growth, total amounts of MFA and SFA (76%) and linolenic acid C18:3.
and minimal in Botryococcus sp. A1162 (22.1%). Nevertheless, it is worth noting the presence of hexadecatetraenoic
The highest total content of SFA in all strains during nitrogen acid С16:4 (3.9 ± 0.5%) during nitrogen depletion in a
starvation was found in C. vulgaris A1123, S. abundans A1175 and biomass of S. abundans A1175 and also some minor amounts of
C. sorokiniana A1135, which had levels of 47.6, 51.3 and 46.7% of the octadecatetraenoic acid С18:4 (0.3 ± 0.0%), undetected by the re-
total fatty acid content, respectively. These strains also had very similar sults GC–MS analysis in study [29].
total amounts of unsaturated fatty acids (UFA): 52.4, 48.7 and 53.3%, re-
spectively. Nevertheless, during nitrogen depletion, the biggest total
amounts of SFA and MFA, accompanied by a high specific growth rate 3.7. Protein and carbohydrate content
in nitrogen replete conditions, were found in C. vulgaris A1123 (67.0%)
and S. abundans A1175 (72.8%). Another feature of the S. abundans Microalgae are also considered as a valuable stock for the industry,
A1175 was a relatively constant level of production of SFA, MFA and as some strains have high protein and carbohydrate content [30,31].
PUFA throughout the cultivation, while the majority of other strains Therefore, to evaluate the applicability of the strains C. vulgaris A-1123
had significant differences in the number of saturated bonds in the and S. abundans A1175 for the production of protein and carbohydrates
fatty acids, depending on growth phase. S. abundans A1175, when as a byproduct of their biomass processing, we have performed

Table 1
Analysis of the properties of microalgal strains under study.

Strain μ, d−1 DT Biomass productivity, Lipid productivity, mg L−1 d−1 Lipid content, %
mg L−1 d−1

Chlorella vulgaris A1123 0.09 ± 0.01 8.0 ± 0.7 35.8 ± 0.24 13.7 ± 0.2 38.3 ± 1.0
Scenedesmus abundans A1175 0.20 ± 0.01 3.6 ± 0.5 73.82 ± 4.53 32.8 ± 1.5 44.4 ± 2.7
Chlorella sorokiniana A1135 0.05 ± 0.01 13.5 ± 0.1 10.60 ± 0.34 3.3 ± 0.1 31.2 ± 3.4
Scenedesmus obliquus A1167 0.07 ± 0.01 9.5 ± 1.0 20.85 ± 0.71 8.6 ± 0.7 41.2 ± 2.2
Botryococcus sp. A1162 0.11 ± 0.01 6.1 ± 0.5 29.61 ± 1.46 9.9 ± 0.3 33.4 ± 2.3
A.V. Piligaev et al. / Algal Research 12 (2015) 368–376 375

used in [35], which was 51.0 ± 0.7% carbohydrate. At the same time,
S. abundans A1175 had a carbohydrate content of 43.3 ± 1.2%, which
exceeded the corresponding data for Scenedesmus sp. CCNM 1077 [34].
Thus, based on these data, C. vulgaris A-1123, during nitrogen starvation,
exceeds S. abundans A1175 in protein content but has a comparable
amounts of carbohydrates. Thus, both of these strains can be used as a
source of carbohydrates, which can then be converted into valuable
products, e.g., ethanol [36].

4. Conclusions

In this work, we have studied the properties of a number of strains


isolated from different climatic conditions, all of which were easily
maintained as a pure cultures in our laboratory conditions. Isolated
strains belonged to genera Chlorella, Botryococcus and Scenedesmus, all
of which are widespread in nature. It was shown that the microalgal
strains have different fatty acid compositions in various stages of
Fig. 7. TLC analysis of the lipid extracts of microalgal strains in exponential and stationary
growth phases. A — exponential phase of growth; B — stationary phase of growth; 1 and 6
growth, and, generally, the amount of SFA decreased in the stationary
— Chlorella vulgaris A1123; 2 and 7 — Scenedesmus abundans A1175; 3 and 8 — Chlorella phase, excluding strain S. abundans A1175, which had a relatively con-
sorokiniana A1135; 4 and 9 —Scenedesmus obliquus A1167; 5 and 10 — Botryococcus sp. stant composition of SFA, MFA and PUFA. It was shown that the proper-
A1162. ties of S. abundans A1175 significantly surpassed the properties of other
strains of microalgae isolated in this work, with regard to biofuel appli-
cations. This strain has a high growth rate and high lipid (including tri-
additional analysis of biomass composition, namely the content of car-
glycerides) content by biomass (44.4 ± 2.7%) with 72.8% SFA and MFA.
bohydrates and protein in nitrogen depleted conditions.
It is worth noting that the use of S. abundans A1175 for biofuel produc-
It has been shown that the protein content in the biomass of strain C.
tion is preferable because of its relatively low content of PUFA, which is
vulgaris A-1123 in the stationary phase of growth was 26.0 ± 1.2%. This
strongly regulated by biofuel standards. However, another strain, C.
is lower than the other strains. For example, C. vulgaris [32] contains
vulgaris A1123, also had a high content of SFA and MFA (67.0%), but it
47.82 ± 0.05% of protein. Protein content of S. abundans A1175 was
also had a specific growth rate of two times less than S. abundans
15.7 ± 0.7%, which was significantly lower than other closely related
A1175. C. vulgaris A1123 had higher protein content during nitrogen
species, e.g. Scenedesmus sp. NT1d (33.08%) and Scenedesmus dimorphus
starvation than S. abundans A1175, while the amounts of carbohydrate
NT8c (22.66%) [33]. In addition, in the work [34], it has been shown that
content was comparable in both strains. Thus, S. abundans A1175 is a
the complete absence of nitrogen in the medium results in reduced pro-
promising strain for producing high-quality biodiesel that meets mod-
duction of protein by microalgae Scenedesmus sp. CCNM 1077, from a
ern industrial standards.
biomass percentage of 47.75 to 16.87. Thus, the low protein content
compared with other strains may be explained by their individual char-
acteristics under nitrogen starvation. Acknowledgment
The carbohydrate content in the stationary phase for C. vulgaris A-
1123 was 37.8 ± 1.4%. This is higher than that in C. vulgaris [32], This work was done by the financial support of the Russian
which was 8.08 ± 0.09% carbohydrate, but less than C. vulgaris FSP-E Foundation for Basic Research (project no. 14-08-31589 mol_a).

Table 2
FAME analysis with GC–MSa. Given values are expressed in % as mean ± standard deviation.*a

Scenedesmus abundans
Chlorella vulgaris A1123 Chlorella sorokiniana Scenedesmus obliquus Botryococcus sp. A1162
FA A1175
A1135 A1167

Exp St Exp St Exp St Exp St Exp St

C14:0 1.2 ± 0.1 1.1 ± 0.4 1.0 ± 0.1 1.5 ± 0.2 1.2 ± 0 1.2 ± 0.2 1.2 ± 0.2 1.1 ± 0.1 1.5 ± 0.1 1.4 ± 0
C16:0 18.7 ± 1.6 18.6 ± 1.8 17.5 ± 0.5 22.0 ± 5.1 16.1 ± 0.6 16.0 ± 3.1 17.2 ± 0.5 14.3 ± 0.9 16.0 ± 0.6 14.4 ± 0.7
C16:1 1.0 ± 0.2 1.3 ± 0.1 1.5 ± 0.3 1.5 ± 0.3 2.1 ± 0.5 2.4 ± 0.2 1.8 ± 0.2 2.8 ± 0.2 3.8 ± 0.3 4.5 ± 0.9
C16:2 3.1 ± 0.7 5.6 ± 0.4 1.3 ± 0.1 2.2 ± 0.4 3.2 ± 0.8 4.0 ± 0.4 0.7 ± 0.1 0.7 ± 0.1 7.0 ± 0.5 8.9 ± 0.1
C16:3 1.1 ± 0.2 1.7 ± 0.3 1.6 ± 0.0 2.3 ± 0.4 7.8 ± 0.9 6.8 ± 0.3 0.8 ± 0.1 1.3 ± 0.0 0.9 ± 0.1 1.4 ± 0.3
C16:4 0.5 ± 0.6 0.5 ± 0.0 3.6 ± 0.7 3.9 ± 0.5 0.8 ± 0.2 0.6 ± 0.1 7.5 ± 1.7 10.5 ± 0.6 2.1 ± 0.1 2.5 ± 0.1
C18:0 42.3 ± 3.8 27.8 ± 4.9 36.5 ± 4.8 27.5 ± 0.1 35.6 ± 5.3 29.3 ± 4.6 33.8 ± 4.6 19.7 ± 0.7 32.1 ± 1.6 20.2 ± 1.5
C18:1 20.0 ± 4.6 17.7 ± 0.6 23 ± 3.5 18.2 ± 3.6 13.2 ± 0.5 16.8 ± 2.7 16.6 ± 1.4 18.3 ± 4.5 18.3 ± 3.9 18.4 ± 0.9
C18:2 9.5 ± 3.1 21.5 ± 3.4 8.0 ± 1.6 11.8 ± 2.4 6.0 ± 0.7 8.2 ± 1.8 3.9 ± 0.6 6.9 ± 0.8 13.1 ± 0.2 20.2 ± 4.7
C18:3 2.3 ± 0.5 3.3 ± 0.2 6.1 ± 1.4 6.6 ± 1.1 14.0 ± 4.0 13.8 ± 0.8 13.4 ± 3.2 20.5 ± 2.9 5.3 ± 1.2 6.6 ± 1.6
C18:4 – 0.4 ± 0.0 – 0.3 ± 0.0 – – 1.4 ± 0.8 2.5 ± 0.4 – 0.9 ± 0.2
C20:0 – 0.1 ± 0.0 – 0.3 ± 0.0 – 0.2 ± 0.0 – – – –
C20:1 0.3 ± 0.0 0.5 ± 0.1 – 1.8 ± 0.1 – 0.8 ± 0.1 1.7 ± 0.5 1.2 ± 0.2 – 0.7 ± 0.1
SFA 62.2 47.6 55.0 51.3 52.9 46.7 52.2 35.2 49.6 36.0
UFA 37.8 52.4 45.0 48.7 47.1 53.3 47.8 64.8 50.4 64.0
MFA 21.3 19.5 24.5 21.5 15.3 20.0 20.0 22.3 22.0 23.5
PUFA 23.1 33.0 25.0 27.1 26.4 33.4 23.7 23.7 42.4 22.1
SFA +
83.5 67.0 79.4 72.8 68.3 66.6 72.2 57.7 71.7 59.6
MFA
a
Exp — exponential phase of growth; St — stationary phase of growth; FA — fatty acids; UFA — unsaturated FA
376 A.V. Piligaev et al. / Algal Research 12 (2015) 368–376

References [19] M. Reckermann, Flow sorting in aquatic ecology, Sci. Mar. 64 (2000) 235–246.
[20] M. Cellamare, A. Rolland, S. Jacquet, Flow cytometry sorting of freshwater phyto-
[1] H.V. Thurman, Introductory Oceanography, Prentice Hall College, New Jersey, USA, plankton, J. Appl. Phycol. 22 (2010) 87–100.
1997. [21] N. Hempel, I. Petrick, F. Behrendt, Biomass productivity and productivity of fatty
[2] K.N. Sorokina, V.A. Yakovlev, A.V. Piligaev, R.G. Kukushkin, S.E. Pel'tek, N.A. acids and amino acids of microalgal strains as key characteristics of suitability for
Kolchanov, V.N. Parmon, Potential of microalgae as a source of bioenergy, Catal. biodiesel production, J. Appl. Phycol. 24 (2012) 1407–1418.
Ind. 4 (2012) 202–208. [22] G. Knothe, Dependence of biodiesel fuel properties on the structure of fatty acid
[3] Y. Chisti, Biodiesel from microalgae, Biotechnol. Adv. 25 (2007) 294–306. alkyl esters, Fuel Process. Technol. 86 (2005) 1059–1070.
[4] T.M. Mata, A.A. Martins, N.S. Caetano, Microalgae for biodiesel production and other [23] G. Knothe, J. Van Gerpen, J. Krahl, The Biodiesel Handbook, AOCS Press, Champaign,
applications: a review, Renew. Sust. Energ. Rev. 14 (2010) 217–232. Illinois, 2005.
[5] J. Singh, S. Cu, Commercialization potential of microalgae for biofuels production, [24] G. Knothe, "Designer" biodiesel: optimizing fatty ester (composition to improve fuel
Renew. Sust. Energ. Rev. 14 (2010) 2596–2610. properties, Energy Fuel 22 (2008) 1358–1364.
[6] P.M. Schenk, S.R. Thomas-Hall, E. Stephens, U.C. Marx, J.H. Mussgnug, C. Posten, O. [25] M.S. Graboski, R.L. McCormick, Combustion of fat and vegetable oil derived fuels in
Kruse, B. Hankamer, Second generation biofuels: high-efficiency microalgae for bio- diesel engines, Prog. Energy Combust. Sci. 24 (1998) 125–164.
diesel production, BioEnergy Res. 1 (2008) 20–43. [26] R.A.I. Abou-Shanab, I.A. Matter, S.N. Kim, Y.K. Oh, J. Choi, B.H. Jeon, Characterization
[7] E.C. Francisco, D.B. Neves, E. Jacob-Lopes, T.T. Franco, Microalgae as feedstock for and identification of lipid-producing microalgae species isolated from a freshwater
biodiesel production: carbon dioxide sequestration, lipid production and biofuel lake, Biomass Bioenergy 35 (2011) 3079–3085.
quality, J. Chem. Technol. Biotechnol. 85 (2010) 395–403. [27] D.K.Y. Lim, S. Garg, M. Timmins, E.S.B. Zhang, S.R. Thomas-Hall, H. Schuhmann, Y. Li,
[8] K.-L. Yeh, J.-S. Chang, Effects of cultivation conditions and media composition on cell P.M. Schenk, Isolation and evaluation of oil-producing microalgae from subtropical
growth and lipid productivity of indigenous microalga Chlorella vulgaris ESP-31, coastal and brackish waters, PLoS One 7 (2012).
Bioresour. Technol. 105 (2012) 120–127. [28] Y.N. Liang, N. Sarkany, Y. Cui, Biomass and lipid productivities of Chlorella vulgaris
[9] J.R. Stein, Handbook of Phycological Methods Culture Methods and Growth under autotrophic, heterotrophic and mixotrophic growth conditions, Biotechnol.
Measurements, Cambridge University Press, 1980. Lett. 31 (2009) 1043–1049.
[10] H. Pereira, L. Barreira, A. Mozes, C. Florindo, C. Polo, C. Duarte, L. Custódio, J. Varela, [29] S.K. Mandotra, P. Kumar, M.R. Suseela, P.W. Ramteke, Fresh water green microalga
Microplate-based high throughput screening procedure for the isolation of lipid- Scenedesmus abundans: a potential feedstock for high quality biodiesel production,
rich marine microalgae, Biotechnol. Biofuels 4 (2011) 1–12. Bioresour. Technol. 156 (2014) 42–47.
[11] S.F. Altschul, T.L. Madden, A.A. Schaffer, J.H. Zhang, Z. Zhang, W. Miller, D.J. Lipman, [30] E.W. Becker, Micro-algae as a source of protein, Biotechnol. Adv. 25 (2007) 207–210.
Gapped BLAST and PSI-BLAST: a new generation of protein database search pro- [31] E.W. Becker, Microalgae in Human and Animal Nutrition, in: A. Richmond (Ed.),
grams, Nucleic Acids Res. 25 (1997) 3389–3402. Handbook of Microalgal Culture: Biotechnology and Applied Phycology, Blackwell,
[12] K. Tamura, D. Peterson, N. Peterson, G. Stecher, M. Nei, S. Kumar, MEGA5: molecular Oxford, 2004.
evolutionary genetics analysis using maximum likelihood, evolutionary distance, [32] O. Tokusoglu, M.K. Unal, Biomass nutrient profiles of three microalgae: Spirulina
and maximum parsimony methods, Mol. Biol. Evol. 28 (2011) 2731–2739. platensis, Chlorella vulgaris, and Isochrisis galbana, J. Food Sci. 68 (2003) 1144–1148.
[13] N. Saitou, M. Nei, The neighbor-joining method — a new method for reconstructing [33] V.T. Duong, F. Ahmed, S.R. Thomas-Hall, K. Nowak, P.M. Schenk, High protein- and
phylogenetic trees, Mol. Biol. Evol. 4 (1987) 406–425. high lipid-producing microalgae from outback Australia as potential feedstock for
[14] J. Felsenstein, Confidence-limits on phylogenies — an approach using the bootstrap, animal feed and biodiesel, Front. Bioeng. Biotechnol. 3 (2015).
Evolution 39 (1985) 783–791. [34] I. Pancha, K. Chokshi, B. George, T. Ghosh, C. Paliwal, R. Maurya, S. Mishra, Nitrogen
[15] Y. Collos, F. Mornet, A. Sciandra, N. Waser, A. Larson, P.J. Harrison, An optical method stress triggered biochemical and morphological changes in the microalgae
for the rapid measurement of micromolar concentrations of nitrate in marine phy- Scenedesmus sp. CCNM 1077, Bioresour. Technol. 156 (2014) 146–154.
toplankton cultures, J. Appl. Phycol. 11 (1999) 179–184. [35] S.H. Ho, S.W. Huang, C.Y. Chen, T. Hasunuma, A. Kondo, J.S. Chang, Characterization
[16] D. Pal, I. Khozin-Goldberg, Z. Cohen, S. Boussiba, The effect of light, salinity, and ni- and optimization of carbohydrate production from an indigenous microalga Chlorel-
trogen availability on lipid production by Nannochloropsis sp, Appl. Microbiol. la vulgaris FSP-E, Bioresour. Technol. 135 (2013) 157–165.
Biotechnol. 90 (2011) 1429–1441. [36] H.-W. Yen, I.C. Hu, C.-Y. Chen, S.-H. Ho, D.-J. Lee, J.-S. Chang, Microalgae-based
[17] M. Dubois, K.A. Gilles, J.K. Hamilton, P.A. Robers, F. Smith, Colorimetric method for biorefinery — from biofuels to natural products, Bioresour. Technol. 135 (2013)
determination of sugars and related substances, Anal. Chem. 28 (1956) 350–356. 166–174.
[18] T.Y.D. Thi, B. Sivaloganathan, J.P. Obbard, Screening of marine microalgae for biodie-
sel feedstock, Biomass Bioenergy 35 (2011) 2534–2544.

Вам также может понравиться