Open access peer-reviewed chapter

The Colonial Microalgae Botryococcus braunii as Biorefinery

Written By

Edmundo Lozoya-Gloria, Xochitl Morales-de la Cruz and Takehiro A. Ozawa-Uyeda

Submitted: 21 June 2019 Reviewed: 24 June 2019 Published: 21 September 2019

DOI: 10.5772/intechopen.88206

From the Edited Volume

Microalgae - From Physiology to Application

Edited by Milada Vítová

Chapter metrics overview

1,274 Chapter Downloads

View Full Metrics

Abstract

The growing shortage of fossil fuels caused an increase in the demand for alternative and renewable fuels. Biofuels, like bioethanol and biodiesel, have received more attention as a sustainable replacement of fossil fuels. However, these have a poor oxidative stability, little energy content by volume, and many oxygenated compounds, which may cause corrosion and damage to the engines. Therefore, they are used as a mixture with standard fuels. Some species of microalgae are candidates to produce oils as triglycerides (TGA) to produce biodiesel by transesterification; however, the problem will remain. The colonial microalgae Botryococcus braunii produces and accumulates a high amount of long-chain nonoxygenated hydrocarbons, similar to those obtained from the fractionated distillation of crude petroleum. This is one of the few organisms reported to have a direct contribution in the formation of the oil reserves currently in use. Additionally, B. braunii produces pigments and long-chain carbohydrates that have interesting properties for various industries. There are still problems to be solved in order to consider it as economically viable and profitable, but important progress is being made. Therefore, this microalga is very attractive for the synthesis of hydrocarbons and other value-added compounds, making it an interesting biorefinery organism.

Keywords

  • biorefinery
  • Botryococcus
  • exopolysaccharides
  • hydrocarbons
  • lipids
  • pigments

1. Introduction

Botryococcus braunii is a colonial microalga Trebouxiophyceae, distributed in brackish and sweet water [1]. It reaches densities of 1.4 × 106 colonies/L [2], and its geochemistry significance is important. Paleobotanical studies suggest that it is one of the largest sources of hydrocarbons in oil-rich deposits dating back to the Ordovician period [1, 3, 4, 5]. It is the only colonial microalga that accumulates and secrets liquid hydrocarbons (Figure 1), and depending on the strain and growing conditions, race B can accumulate hydrocarbons up to 85% and race A up to 61% of their dry weight.

Figure 1.

B. braunii race B colony secreting liquid hydrocarbons.

B. braunii is related with Characium vaculatum and Dunaliella parva [1]. Due to the hydrocarbons and the molecular phylogeny of B. braunii [6], it is classified in three races (A, B, and L). Race A produces n-alkadienes and alkatrienes of C23–C33 [7], although two unusual hydrocarbons have been characterized, the triene C27H51 and tetraene C27H48 [1]. Race A hydrocarbon dry weight varies from 0.4 to 61% [7, 8]. Race B produces triterpenoids hydrocarbons known as botryococcenes (CnH2n−10, n = 30-37) [9] and methylsqualenes C31–C34 [10, 11]. The botryococcenes can be from 27 to 86% of the dry weight [12]. Race L produces a tetraterpene C40 known as lycopadiene and constitutes from 0.1 to 8% of the dry weight [13, 14]. This race contains 5% of lycopatriene, lycopatetraene, lycopapentaene, and lycopahexaene [15]. In addition, a race S is proposed, which synthesizes saturated n-alkanes C18 and C20, and epoxy-alkanes; however, its existence is not yet fully accepted [6].

After the hydrocracking process and subsequent distillation, race B hydrocarbons become biofuels currently used in internal combustion engines [16] as shown in Figure 2.

Figure 2.

Hydrocarbons produced by the B. braunii races. Biofuels derived from race B are shown. RON, research octane number = 92–98, this is a measure of autoignition resistance in a spark-ignition engine. In the USA: regular (97 RON) and premium (95 RON). Adapted from [16, 17, 18].

Advertisement

2. Physiology and biochemistry of Botryococcus braunii

B. braunii races differ also by its morphological and physiological characteristics. Cells from A and B races are of 13 μm × 7–9 μm, and those of L race are 8–9 μm × 5 μm [19].

Each colony is constituted by a group of 50–100 piriform cells embedded in a hydrocarbon network and the extracellular matrix (ECM). This ECM contains three main components: (1) a fibrous cell wall surrounding each cell and having β-1,4-and/or β-1,3-glucans including cellulose; (2) the intracolonial space constituted by a network of liquid hydrocarbons; and (3) a fibrillary sheath composed mainly of arabinose and galactose polysaccharides, holding the liquid hydrocarbons [20].

B. braunii may have a hetero-, mixo-, or phototrophic grow and the morphology will depend on the C source and the amount of light [21]. The hydrocarbon production is associated with the cell division [22], likely due to the localization of the enzymes involved in the alkadienes, alkatrienes (race A), and botryococcenes (race B) biosynthesis [23].

Other difference among the races is the keto-carotenoid accumulation in the stationary phase of cultures. Races B and L change color from green-brown to orange, and race A changes from green to yellow-orange [1]. The production of carotenoids is also a stress response by environmental factors. The DAD1 gene expression, a suppressor of programmed cell death, was reported in race B, under stress conditions at 10–60 min [24]. B. braunii is tolerant to desiccation and extreme temperatures, which allows its global dispersion in different environments [25]. The reproduction mechanism of B. braunii seems to be autosporic [26].

Symbiotic bacteria have been reported after microscopic observations, and an ectosymbiont α-proteobacteria (BOTRYCO-2) that promotes the productivity of biomass and hydrocarbons was described [2, 27].

2.1 Biosynthesis of alkadienes and alkatrienes

Characteristic alkadienes and alkatrienes of race A have double links and similar stereochemistry as oleic acid. Experiments with labeled fatty acids have shown that this one is the main precursor by the long-chain fatty acids (LCFAs) pathway, followed by a decarboxylation process [1, 17, 28, 29]. The first step is the elongation of oleic acid (18:1 cis-Δ9) and its isomer elaidic acid (18:1 trans-Δ9). The acyl-CoA reductase and decarbonylase enzymes in race A microsomes suggest an alternative mechanism where the LCFAs are reduced to aldehydes and decarbonylated to produce alkadienes and alkatrienes [17, 30]. Race A transcriptome allowed the identification of six candidate genes potentially involved in this biosynthesis [31].

2.2 Biosynthesis of botryococcenes

The analysis of race B transcriptome and other evidences suggests that the biosynthesis of isoprenoids comes from the deoxyxylulose phosphate/methylerythritol phosphate (DXP/MEP) pathway [32, 33, 34]. Expressed sequence tag (EST) markers for enzymes of the DXP/MEP pathway [34], as well as multiple isoforms of enzymes for the 3-phospho-d-glycerate biosynthesis from d-glyceraldehyde-3-phosphate and pyruvate as precursors, were identified. Some of the respective transcripts are present in high abundance (>250 reads/Kb), suggesting a high metabolic flow in B. braunii [31].

The first step is the formation of 1-deoxy-d-xylulose-5-phosphate (DOXP) by the DOXP synthase (DXS) (Figure 3).

Figure 3.

Biosynthesis of tri- and tetraterpenes in B. braunii race B. (a) FPP production; (b) carotenoid production from GGPP; (c) squalene production from FPP; (d) methylated botryococcene production; (e) methylated squalene production. BSS, Botryococcus squalene synthase; CtrB, phytoene-synthase; DXR, 1-deoxy-d-xylulose-5-phosphate reductase; DXS, 1-deoxy-d-xylulose-5-phosphate synthase; FPPS, farnesyl diphosphate synthase; GPPS, geranyl diphosphate synthase; NADPH+ and NADP+, nicotinamide adenine dinucleotide phosphate (reduced and oxidized); PPi, inorganic pyrophosphate; PSPP, cyclopropyl presqualene diphosphate; SAM, S-adenosyl methionine; SAH, S-adenosyl-l-homocysteine; SSL, squalene synthase-like; SMT, squalene methyltransferase; TMT, triterpene methyltransferase. Adapted from [17, 34].

The characterization of three DXS isoenzymes in race B shows that they are active and have similar kinetic parameters, which increases the metabolic flow for the production of terpenoids [35]. The DOXP is reduced by the DXP reductoisomerase (DXR) to 2-C-methylerythritol-4-phosphate (MEP), and converted to isopentenyl diphosphate (IPP) and dimethylallyl pyrophosphate (DMAPP). In the B. braunii transcriptome, only one DXR has been found [34]. The next step involves condensation of IPP and DMAPP to form geranyl diphosphate (GPP), and the addition of other IPP produces farnesyl diphosphate (FPP) [17] (Figure 3a). Two B. braunii genes code for farnesyl diphosphate synthase isoenzymes (FPPS) with an amino acid identity of 72% [34].

Addition of another IPP forms the geranylgeranyl diphosphate (GGPP), precursor of the tetraterpenoid carotenoids (Figure 3b). This begins with the formation of a trans-isoprenyl diphosphate by the phytoene synthase (CtrB) enzyme, condensing two GGPP molecules in two steps with the release of pyrophosphate. In the first step, (1R, 2R, 3R)-prephytoene diphosphate is produced from half cyclopropyl (C1′-2-3) reordered to provide 15-cis-phytoene, which can be converted into a wide variety of carotenoids [34, 36, 37, 38]. All are important antioxidant photoprotectors and modulators of the function of membrane proteins for photosynthetic complexes [39].

The squalene production [40] starts with the Botryococcus squalene synthase (BSS) enzyme, using two FPP molecules. Botryococcenes production uses also two FPP molecules but the product is the intermediary cyclopropyl presqualene diphosphate (PSPP) (Figure 3c). With NADPH, the PSPP has two options; one forms the botryococcene with a C3-C1 connection between the FPP molecules (Figure 3d). The other option forms a C1-C1′ between two FPP molecules producing squalene that will be methylated (Figure 3e) further on. These reactions are catalyzed by squalene synthase-like (SSL) enzymes. Three SSL genes have been identified but none is directly related with the botryococcene biosynthesis [41]. However, when the 3 SSLs enzymes were mixed in vivo and in vitro, botryococcene (SSL-1 + SSL-3) or squalene (SSL1 + SSL-2) was synthesized. SSL-1 condenses two FPP molecules to produce PSPP [42], demonstrating the versatility and potential for metabolic engineering of botryococcene biosynthesis.

Most botryococcenes are excreted to the ECM where they are methylated. The di- and tetramethyl forms are related to six genes coding for triterpene and squalene methyltransferases (TMT, SMT) [43] (Figures 3d and 3e). The botryococcenes are methylated to produce C31–C37 hydrocarbons, C34 being the main in race B. Three cyclic botryococcene C33 molecules and a trimethylsqualene isomer were recently found [44]. Also, two squalene epoxidase (BbSQE-I and -II) enzymes converting squalene into membrane sterols were identified [45]. Data of the B. braunii race B nuclear genome will allow the search for possible regulatory routes of this singular metabolism [46].

2.3 Biosynthesis of lycopadiene

The formation of lycopadiene of race L is similar to the squalene. In the B. braunii transcriptome, there are two homologous contigs to squalene synthase (SS) [31]. One encodes a squalene synthase (LSS) and the other for a lycopaoctaene synthase (LOS). LOS uses preferentially in vivo GGPP, and C15 and C20 prenyl diphosphates as substrates [15] (Figure 4).

Figure 4.

Lycopadiene biosynthetic pathway. (a) Reduction of GGPP to PPP and condensation by LOS. (b) LOS condensation of GGPP to form phytyl diphosphate and reduction to lycopaoctaene. (c) FPP use by LSS or LOS for squalene production. DXR, 1-deoxy-d-xylulose-5-phosphate reductase; DXS, 1-deoxy-d-xylulose-5-phosphate synthase; FPP, farnesyl diphosphate; FPPS, farnesyl diphosphate synthase; GGPP, geranylgeranyl diphosphate; GPPS, geranyl diphosphate synthase; GPPR, geranyl diphosphate reductase; NADPH+ and NADP+, nicotinamide adenine dinucleotide phosphate (reduced and oxidized); PPi, inorganic pyrophosphate; PPP, phytyl diphosphate; PLPP, prelycopaoctaene diphosphate; LOS, lycopaoctaene synthase; LSS, B. braunii race L squalene synthase. Adapted from [15].

There are two biosynthetic mechanisms for lycopadiene from C20 prenyl diphosphate intermediates. In one, the GGPP reduction by a GGPP-reductase produces phytyl diphosphate (PPP), and LOS condenses two PPP molecules producing lycopadiene (Figure 4a). In the other one, LOS condenses two GGPP molecules producing prelycopaoctaene diphosphate (PLPP), which rearranges into lycopaoctaene. Finally, lycopadiene seems to be produced by enzymatic reductions not yet identified (Figure 4b).

LOS may also form squalene from FPP (Figure 4c). These results show the plasticity of L race to synthesize squalene and lycopadiene.

2.4 Extracellular matrix (ECM) polymers

ECM contains long chains of polymerized polyacetal hydrocarbons joined to specific hydrocarbons of each race. There is a fibrillary sheath that envelops the entire colony, formed mainly by arabinose (42%) and galactose (39%). The cell wall contains β-1,4 and/or β-1,3 glucans making a cellulose-like polymer [20].

Also, there's a biopolymer resistant to nonoxidative chemical degradation as acetolysis. This biopolymer resembles sporopollenins [1] of the outer walls of pollen grains and spores of microorganisms [47]. It seems to be formed by oxidized carotenoid polymers and phenolic compounds that absorb UV-B light as p-coumaric and p-ferulic acids [48].

Advertisement

3. Profitability of B. braunii derivatives

3.1 Hydrocarbons

Both bioethanol and biodiesel have a poor oxidative stability, low energy content by volume, and high content of oxygenated compounds, which damage combustion engines and cause corrosion, erosion, and accumulation of deposits in the nozzles; because of these reasons, they are mixed with standard fuels [49, 50]. B. braunii accumulates hydrocarbons similar to those of the crude oil, and their direct contribution in the formation of oil reserves currently in use has been reported [3, 4, 5]. The B. braunii oils showed almost equal values in density and surface tension than the diesel, but with higher kinematic viscosity and distillation temperature [50]. The B. braunii race B oil was already converted into diesel with an 85% performance, using a simple conversion process under mild conditions of 260°C and 1 atm. The physical properties are relatively close to the specification for diesel, with 40 as estimated cetane (CN) number [51].

The limitation to use B. braunii as biorefinery is the slow growth rate of days in comparison with hours in other algae [49, 52]. Other factors affecting the growth and hydrocarbon production are the strain, CO2, light, water, nutrients, temperature, pH, and salinity [53, 54, 55, 60] (Table 1). A JET PASTER treatment was used to do a mechanical cell disruption and removal of the polysaccharides of the B. braunii colonies, increasing the hydrocarbon extraction up to 82.8%. This treatment did not affect the photosynthetic function of the cells [56]. On the other hand, a repetitive nondestructive extraction with heptane was reported as having some advantages [57]. Also, a continuous growth and extraction column of n-dodecane was reported recently as an efficient hydrocarbon extraction method without significant loss of the viability of the cells [58]. Considering these milking procedures and achieving a 10% rate of return, a minimum sales price (MSP) of US$3.20 per liter was calculated, and a reduction down to US$1.45 per liter was proposed, if hydrocarbon content increases and extraction procedures become more efficient [59].

St Culture conditions SCGR Dt THC Ref.
°C PAR Php CO2
Showa (B) 30 850 14:10 1 0.5 1.40 NIA [54]
Showa (B) 25, 30 85–398 14:10 1.0–10.0 0.19–0.44 1.60–3.60 30–39 [54]
Showa (B) 23–25 250 24 0.3 0.42 1.70 24–29 [52]
Showa (B) 23 150 16:8 2 0.17 4.08d 25 [55]
Yayoi (B) 25 240 12:12 2 0.2 3.50 40.5 [38]
AC759 (B) 23 150 16:8 2 0.07 9.90d 21 [55]
AC761 (B) 23 150 16:8 2 0.11 6.30d 45 [55]
IPE001 (B) 25 35 16:8 1 0.15c 4.50c 64.3 [61]
BOT-144 (B) 25 60a 24 0 0.16 4.33d 50 [62]
LB-572 (A) 26 12 Klux 24 2 0.07c 10.60c 28 [53]
Gottingen 807/1 (A) 25 26b 14:10 1 0.3 2.30 40.5 [67]
AC755 (A) 23 150 16:8 2 0.05 13.86d 16 [55]
CCALA777 (A) 23 150 16:8 2 0.06 11.55d 10 [55]
CCALA778 (A) 23 150 16:8 2 0.17 4.08d 0 [55]
CCAp807/2 (A) 23 150 16:8 2 0.11 6.30d 7 [55]
765 25 150 24 20 0.13c 5.50c 24 [64]
765 25 120 24 ASLW NIA NIA 23.8 [65]
GUBIOTJTBB1 25 35 16:8 0 0.112 6.19 52.6 [66]
AP 103 23 30 16:8 0 NIA NIA 13 [67]

Table 1.

Comparison of culture conditions and productivity of hydrocarbons between B. braunii strains at laboratory scale.

Blue light λ = 470 nm.


W/m2.


Estimated values [54].


Calculated values from μ, using Ln(2)/μ equation.


ASLW, aerated swine lagoon wastewaters (not sterile); °C, temperature; CO2, % v/v; Dt, doubling time (days); NIA, no information available; PAR, photosynthetic active radiation (μmols of photons/m2 s); Php, photoperiod (light/dark hours); SCGR, specific cell growth rate (μ/day); μ, specific velocity of growth rate; St, strain (race); THC, total hydrocarbons (% DW, dry weight).

There are different open and closed culture systems in photobioreactors (PBR) [63, 64], but more studies are required at pilot and industrial scale, to reduce problems by contamination and low yield of biomass and hydrocarbon production [49]. Table 2 summarizes some data about cell growth and hydrocarbon productivity using different culture systems.

St System Cultures Biomass HCs Ref
°C PAR CO2 SCGR Xmax Px CNT WHC
GUBIOT JTBB1 Plain (3 L) 25 35 (16 h) 0% 0.112 NIA 13 52.6 6.8 [62]
765 Column (3 L) 25 150 (24 h) 20% 0.13g NIA 92.4 24.45g 22.6 [64]
Showa (B) PBRa 25-28 282 (15 h) 5–7% NIA 20 1500 22.5 225-340 [68]
NIA PBRb 25 270 (24 h) Mixo-trophic NIA 4.55 234 29.7 71.1 [69]
UTEX-LB 572 (A) Circular (50 L) rT Sol r 0% NIA NIA 77.8 19 13.2 [70]
N-836 (B) Rcwy (80 L) rT Sol r 0% NIA NIA 40 24 10.8 [70]
LB572 (B) PBRc 20 Sol r 0% 0.04 0.3 15 NIA 2.4 [71]
AP103 Rcwy (1800 L) 29 Sol r 5 kWh/m2.day 0% 0.38 NIA 114 11 12.5 [67]
UTEX-LB 572 (A) PBRd 25 55 (24 h) 1% NIA 96.4 0.71i NIA NIA [72]
FACHB 357 (B) Attchde 25 500 (24 h) 1% NIA 62h 5.5–6.5i 19.43 1.06i [73]
TN101 Rcwy scf rT Sol r 0% NIA NIA 33.8i 22.6 8.2-13i [74]

Table 2.

Comparison of culture conditions and productivity of hydrocarbons between strains of B. braunii in bioreactors.

“Tickle film” (30.5 × 16.5 in) continuous.


"Airlift" (10 L).


Panel (1000 L) outdoor and semicontinuous.


"Biofilm" (0.275 m2 or 600 mL).


“Attached” bioreactor (0.08 m2 or 240 mL).


(25 m2 or 5000 L) semicontinuous.


Estimated values [64].


g/m2.


g/m2/day; shadow area indicates the highest reported values up to now.


°C, temperature; CNT, content (% DW dry weight); CO2, % v/v; HCs, hydrocarbons; PAR, photosynthetic active radiation (μmols of photons/m2 s); PBR, photobioreactor; Php, photoperiod (light/dark hours); Px, biomass productivity (mg/L day); NIA, no information available; Rcwy, raceway; rT, room temperature; SCGR, specific cell growth rate (μ/day); μ, specific velocity of growth rate; Sol r, solar radiation; St, strain (race); WHC, weight of hydrocarbons (mg/L day); Xmax, maximum cellular concentration (g/L).

3.2 Lipids

B. braunii also produces saturated and monounsaturated fatty acids, especially palmitic (16:0) and oleic (18:1), as well as triacylglycerols (TAGs). The percentages of total lipids as saturated, monounsaturated, and polyunsaturated fatty acids in dry biomass are around 44.97, 9.85, 79.61, and 10.54%, respectively [64, 75]. Studies in vitro and in vivo showed that these fatty acids effectively improve the absorption of lipophilic drugs like flurbiprofen, through the skin [76].

B. braunii stores TAGs and saturated fatty acids in the lag phase as an adaptation to stress conditions but most are synthesized during the stationary phase. Although highest content of these acids is intracellular, B. braunii secretes oily drops in small quantities observed on the surface of the cell apex [64].

The yield and lipid composition depends on the strain, the culture system used, growth conditions and cell aging, as well as nitrogen, phosphorus, and micronutrient concentrations (Table 3).

St System TRT Biomass Lipids Ref.
SCGR XMax Px CNT Yld. Prod.
UTEX 572 (A) EF (125 mL) 0.04 mM NO3 0.09 0.16 NIA 63 NIA 0.009 [77]
0.37 mM NO3 0.185 0.38 NIA 36 0.19 0.019
KMITL 2 (n.d.) EF (1 L) 86 mg/L NO3 NIA 0.48 NIA 39.42 0.19 NIA [78]
222 mg/L PO4 NIA 0.86 NIA 54.69 0.47 NIA
444 mg/L PO4 NIA 1.91 NIA 23.23a 0.45 NIA
27 mg/L Fe NIA 0.22 NIA 34.93 0.08 NIA
KMITL 2 (n.d.) Outdoor oval pond (150 L) 0.17 g/L NO3 0.045 4.84 NIA 35.24 NIA 0.016 [79]
2.5 g/L NO3 0.049 5.62 NIA 38.60 NIA 0.0189
LB572 (A) FBR column (625 mL) 0083 g/L PO4 and 0.1 g/L SO4 NIA NIA 0.296 64.96 NIA 0.19 [80]
0058 g/L PO4 and 0.09 g/L SO4 NIA NIA 0.304 59.56 NIA 0.18
TRG EF (250 mL) Photoaut. (CO2) 0.093 1.14 NIA 25.1 NIA 0.0241 [81]
Heterot. (gluc 5 g/L) 0.115 1.75 NIA 29.3 NIA 0.0467
Mixot. (gluc 5 g/L + CO2) 0.195 2.46 NIA 37.5 NIA 0.0645
IBL-C117 EF (1 L) Chu (0.75×) 0.13 0.9 0.12 47.1 NIA NIA [82]
Chu (1.0×) 0.13 0.7 0.1 46 NIA NIA
Chu (2.0×) 0.11 1 0.15 41.3 NIA NIA
LB572 (A) EF (1 L) Chu (0.75×) 0.15 1.3 0.18 20.2 NIA NIA [82]
Chu (1.0×) 0.16 1.4 0.2 22.5 NIA NIA
Chu (2.0×) 0.17 1.5 0.22 11 NIA NIA
2441 (A) FBR Airlift (2 L) (N:P = 1:1) in Chu NIA 4.963 0.173 33.7 NIA NIA [83]
(N:P = 3:3) in Chu NIA 3.857 0.215 34.6 NIA NIA
(N:P = 6:6) in Chu NIA 3.987 0.223 32.1 NIA NIA
BOT22 (B) Biofilm bioreac. Nitrocell. Memb. (diam. 25 mm and pore size 0.45 μm) NIA 3.12a 0.42b NIA 0.83a NIA [84]
BOT84 (L) NIA 10.04a 3.8b NIA 1.11a NIA
BOT7 (S) NIA 13.6a 0.99b NIA 0.83a NIA

Table 3.

Comparison of crop conditions and lipid productivity in B. braunii.

mg/cm2.


mg/cm2/day.


CNT, content (% DW dry weight); Chu, Chu media for microalgae [8]; EF, Erlenmeyer flask; FBR, photobioreactor; N:P, proportion of nitrogen: phosphate; Px, biomass productivity (g/L day); NIA, no information available; Prod., productivity (g/L day); Rcwy, raceway; SCGR, specific cell growth rate (μ/day); μ, specific velocity of growth rate; St, strain (race); TRT, treatment; Yld., yield (g/L); Xmax, maximum cellular concentration (g/L).

3.3 Pigments

Algae pigments have been reported to have antioxidant, anticancer, anti-inflammatory, antiobesity, and antiangiogenic properties and function as neuroprotectives [85]. So, they could replace synthetic dyes in food, cosmetic, nutraceutical, and pharmaceutical products [86].

Carotenoid pigments are unsaturated hydrocarbons, while xanthophylls have one or more functional groups containing oxygen such as lutein, canthaxanthin, and astaxanthin [85, 86, 87].

Carotenoids abound in races B and L, lutein being the main pigment (22–29%), followed by others as β-carotene, echinenone, 3-OH echinenone, canthaxanthin, violaxanthin, loroxanthin, and neoxanthin. Transition to stationary phase causes a color change in B. braunii from green to brown, reddish orange, and pale yellow by accumulation of carotenoids and a decrease of intracellular pigments [88]. Canthaxanthin (46%) and echinenone (20–28%) are predominant in the stationary phase in response to nitrogen deficiency [36]. The BOT-20 strain showed a dark red color during growth because of the accumulated echinenone of about 30.5% dry weight and 630 mg/L production, but with few hydrocarbons (8%) [89].

Adonixanthin was detected in race L during the stationary phase [90], and botryoxanthin A, botryoxanthin B, and braunixanthin 1 and 2 were detected in race B [37, 38, 91]. The 2-azahypoxanthine (AHX) similar to the phytohormone induced the accumulation of secondary carotenoids like botryoxanthin A and braunixanthin 1 and decreased the content of botryococcenes during the stationary phase [92], imitating a lack of nitrogen condition without inhibiting the growth.

In race A, lutein (79–84%) is the main carotenoid followed by β-carotene (1.75–2.14%), violaxanthin (6–9%), astaxanthin (3–8%), and zeaxanthin (0.32–0.78%). In salinity and high light intensity conditions, the lutein increases [53, 93]. All of these compounds shown antioxidant properties and inhibitory effect against lipid peroxidation in vitro and in vivo and activated antioxidant enzymes such as catalase [94, 95].

3.4 Polysaccharides

The aqueous extracts of B. braunii (strain LB 572) reduce the skin dehydration, stimulate collagen synthesis, promote the differentiation of adipocytes, and promote antioxidant and anti-inflammatory activities [96]. The extracellular polysaccharides (exopolysaccharides, EPS) constitute most of the organic material of high molecular weight released to the environment by microalgae and other microorganisms. They have antioxidant, immunomodulatory, antibacterial, antiviral, anticarcinogenic, and antihypocholesterolemic effects [97]. They are used as thickeners, emulsifiers, bioflocculants, stabilizers, and gelling agents in foods and cosmetics; are soluble in water; and modify the rheological properties of solutions increasing their viscosity to form gels [1, 98].

The ECM and the fibrillar pod are composed of mucilaginous polysaccharides [20], and other detected EPS are fucose, glucose, mannose, rhamnose, uronic acids, and unusual sugars such as 3-O-methyl fucose, 3-O-methyl rhamnose, and 6-O-methyl hexose [1]. Galactose is involved in the innate and adaptive immune system [99]. l-Arabinose is used as food additive for its sweet taste and poor absorption in humans [100] and is an antiglycemic agent by selective inhibition of invertases, reducing the glycemic response after sucrose ingestion [101]. Uronic acid is a chelating agent to remove metal ions. Fucose has high commercial value for its anticancer properties and for chemical synthesis of flavoring agents [1, 55].

Some B. braunii (UC 58) strains produce 4.0–4.5 g/L EPS with few hydrocarbons (5%). The EPS amount varies with the strain, race, physiological conditions, and culture. Strains of A and B races can produce up to 250 mg/L EPS, and race L up to 1 g/L plus glucose [1].

Greater EPS production correlates with minor growth by N deficiency. Urea and ammonia decrease the pH, as well as EPS production. Optimal conditions for EPS production were nitrate (8 mM) and between 25 and 30°C. Out of these temperatures, the EPS polymerization decreased significantly [1, 102]. Light/dark (16:8) photoperiod produced more hydrocarbons, but continuous light with agitation increased EPS until 1.6 and 0.7 g/L in LB 572 and SAG-30 strains, respectively [103]. EPS production increased (2–3 g/L) in low salinity levels (17–85 mM) as osmoprotectants [53]. High salinity and low N content in D medium induced EPS production (0.549 ± 0.044 g/L) in comparison to the BG11 medium (0.336 ± 0.009 g/L), but biomass was higher in BG11 (1.019 ± 0.051 g/L) than in D (0.953 ± 0.056 g/L) [104]. Modification of culture conditions could be used to increase EPS production, to facilitate the removal, and to increase hydrocarbon recovery. With Botryococcus braunii CCALA 778 (race A), a light:dark cycle at 26°C resulted in an increased production of EPS, and a milking procedure for these polysaccharides has been proposed [105, 106]. EPS can be used as thickening or gelling agents [107].

3.5 Other biopolymers

Algenanes are aliphatic, nonhydrolyzable, and insoluble biopolymers found in the ECM at 9 and 10% dry weight of race A and B, respectively. Due to their high resistance to degradation, they are attributed to the good preservation of colonies in sedimentary rocks [108].

Another reported biopolymer was the polyhydroxybutyrate (PHB), a biodegradable plastic with a yield of about 20% of the dry weight [109]. PHB is a polyester with thermoplastic and biodegradable properties, and it's a carbon and energy storage compound. For its similar physical properties to polypropylene and polystyrene, it is of commercial interest [110]. Under pH 7.5, 40°C, and with 60% wastewater as culture medium, a maximum yield of 247 ± 0.42 mg/L PHB was reported [111].

B. braunii (UTEX 572) was used to produce intra- and extracellular Ag nanoparticles (AgNPs) with antimicrobial properties, and analysis suggested that the exopolysaccharides were the possible reducing and capping agents [112].

Advertisement

4. Conclusions

Although B. braunii has been considered mainly as a good source of biofuels by the possibility to convert its hydrocarbons into currently used fuels, without the necessity of engine modifications, it produces many other high-value derivatives that can be exploited for their promising attractive profits. Besides, along the photosynthetic process, this alga converts 3% of solar energy into hydrocarbons [1] and can reduce CO2 emissions up to 1.5 × 105 tons/year [113]. There are several reports about modifications of the culture conditions through vitamin addition, affecting the yield of several derivatives like biomass, hydrocarbon, and carbohydrate in Botryococcus braunii KMITL 5 [114]; however, those are from not clearly recognized strains and should be carefully taken. With B. braunii race A, B, or L, the main challenge is to accelerate the doubling rate because, depending on the race, it varies between 2 and 10 days. This results in easy contamination with faster growing microorganisms in open ponds used for industrial production, or a high cost of sterile conditions in closed bioreactors. In spite of these disadvantages, we consider that B. braunii is an excellent model of biorefinery. Other strategies to use B. braunii as biorefinery and bioreactor are being developed like the immobilization in polyester [115] or bioharvesting with Aspergillus sp. [116].

Advertisement

Acknowledgments

To the Consejo Nacional de Ciencia y Tecnología (CONACYT) Mexico by the PhD and MSc scholarships of XM-dlC and TAO-U respectively.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Banerjee A, Sharma R, Chisti Y, Banerjee UC. Botryococcus braunii: A renewable source of hydrocarbons and other chemicals. Critical Reviews in Biotechnology. 2002;22:245-279. DOI: 10.1080/07388550290789513
  2. 2. Tanabe Y, Okazaki Y, Yoshida M, Matsuura H, Kai A, Shiratori T, et al. A novel alphaproteobacterial ectosymbiont promotes the growth of the hydrocarbon-rich green alga Botryococcus braunii. Scientific Reports. 2015;5:10467. DOI: 10.1038/srep10467
  3. 3. Moldowan JM, Seifert WK. First discovery of botryococcane in petroleum. Journal of the Chemical Society, Chemical Communications. 1980;19:912-914. DOI: 10.1039/C39800000912
  4. 4. McKirdy DM, Cox RE, Volkman JK, Howell VJ. Botryococcane in a new class of Australian non-marine crude oils. Nature. 1986;320:57-59. DOI: 10.1038/320057a0
  5. 5. Summons RE, Metzger P, Largeau C, Murray AP, Hope JM. Polymethylsqualanes from Botryococcus braunii in lacustrine sediments and crude oils. Organic Geochemistry. 2002;33:99-109. DOI: 10.1016/S0146-6380(01)00147-4
  6. 6. Kawachi M, Tanoi T, Damura M, Kaya K, Watanabe MM. Relationship between hydrocarbons and molecular phylogeny of Botryococcus braunii. Algal Research. 2012;1:114-119. DOI: 10.1016/j.algal.2012.05.003
  7. 7. Metzger P, Berkaloff C, Casadevall E, Coute A. Alkadiene and botryococcene-producing races of wild strains of Botryococcus braunii. Phytochemistry. 1985;24:2305-2312. DOI: 10.1016/S0031-9422(00)83032-0
  8. 8. Metzger P, Largeau C. Botryococcus braunii: A rich source for hydrocarbons and related ether lipids. Applied Microbiology and Biotechnology. 2005;66:486-496. DOI: 10.1007/s00253-004-1779-z
  9. 9. Metzger P, Casadevall E, Pouet MJ, Pouet Y. Structures of some botryococcenes: Branched hydrocarbons from the B-race of the green alga Botryococcus braunii. Phytochemistry. 1985;24:2995-3002. DOI: 10.1016/0031-9422(85)80043-1
  10. 10. Huang Z, Poulter CD. Tetramethylsqualene, a triterpene from Botryococcus braunii var. Showa. Phytochemistry. 1989;28:1467-1470. DOI: 10.1016/S0031-9422(00)97766-5
  11. 11. Achitouv E, Metzger P, Rager MN, Largeau C. C31-C34 methylated squalenes from a Bolivian strain of Botryococcus braunii. Phytochemistry. 2004;65:3159-3165. DOI: 10.1016/j.phytochem.2004.09.015
  12. 12. Brown C, Knights BA, Conway E. Hydrocarbon content and its relationship to physiological state in the green alga Botryococcus braunii. Phytochemistry. 1969;8:543-547. DOI: 10.1016/S0031-9422(00)85397-2
  13. 13. Metzger P, Allard B, Casadevall E, Berkaloff C, Couté A. Structure and chemistry of a new chemical race of Botryococcus braunii Chlorophyceae that produces lycopadiene, a tetraterpenoid hydrocarbon. Journal of Phycology. 1990;26:258-266. DOI: 10.1111/j.0022-3646.1990.00258.x
  14. 14. Metzger P, Pouet Y, Summons R. Chemotaxonomic evidence for the similarity between Botryococcus braunii L race and Botryococcus neglectus. Phytochemistry. 1997;44:1071-1075. DOI: 10.1016/S0031-9422(96)00698-X
  15. 15. Thapa HR, Naik MT, Okada S, Takada K, Molnár I, Xu Y, et al. A squalene synthase-like enzyme initiates production of tetraterpenoid hydrocarbons in Botryococcus braunii Race L. Nature Communications. 2016;7:11198. DOI: 10.1038/ncomms11198
  16. 16. Hillen LW, Pollard G, Wake LV, White N. Hydrocracking of the oils of Botryococcus braunii to transport fuels. Biotechnology and Bioengineering. 1982;24:193-205. DOI: 10.1002/bit.260240116
  17. 17. Cornejo-Corona I, Thapa HR, Devarenne TP, Lozoya-Gloria E. The biofuel potential of the green colonial microalga Botryococcus braunii. In: Torres-Bustillos LG, editor. Microalgae and Other Phototrophic Bacteria. Culture, Processing, Recovery and New Products. 1st ed. New York, USA: Nova Science Publishers, Inc.; 2015. pp. 41-58
  18. 18. Griehl C, Kleinert C, Griehl C, Bieler S. Design of a continuous milking bioreactor for non-destructive hydrocarbon extraction from Botryococcus braunii. Journal of Applied Phycology. 2015;27:1833-1843. DOI: 10.1007/s10811-014-0472-6
  19. 19. Metzger P, Casadevall E, Coute A. Botryococcene distribution in strains of green alga Botryococcus braunii. Phytochemistry. 1988;27:1383-1388. DOI: 10.1016/0031-9422(88)80199-7
  20. 20. Tatli M, Ishihara M, Heiss C, Browne DR, Dangott LJ, Vitha S, et al. Polysaccharide associated protein PSAP from the green microalga Botryococcus braunii is a unique extracellular matrix hydroxyproline-rich glycoprotein. Algal Research. 2018;29:92-103. DOI: 10.1016/j.algal.2017.11.018
  21. 21. Tanoi T, Kawachi M, Watanabe MM. Effects of carbon source on growth and morphology of Botryococcus braunii. Journal of Applied Phycology. 2011;23:25-33. DOI: 10.1007/s10811-010-9528-4
  22. 22. Hirose M, Mukaida F, Okada S, Noguchi T. Active hydrocarbon biosynthesis and accumulation in a green alga, Botryococcus braunii race A. Eukaryotic Cell. 2013;12:1132-1141. DOI: 10.1128/EC.00088-13
  23. 23. Suzuki R, Ito N, Uno Y, Nishii I, Kagiwada S, Okada S, et al. Transformation of lipid bodies related to hydrocarbon accumulation in a green alga, Botryococcus braunii Race B. PLoS One. 2013;8:e81626. DOI: 10.1371/journal.pone.0081626
  24. 24. Cornejo-Corona I, Thapa HR, Browne DR, Devarenne TP, Lozoya-Gloria E. Stress responses of the oil-producing green microalga Botryococcus braunii Race B. Peer J. 2016;4:e2748. DOI: 10.7717/peerj.2748
  25. 25. Demura M, Ioki M, Kawachi M, Nakajima N, Watanabe MM. Desiccation tolerance of Botryococcus braunii Trebouxiophyceae, Chlorophyta and extreme temperature tolerance of dehydrated cells. Journal of Applied Phycology. 2014;26:49-53. DOI: 10.1007/s10811-013-0059-7
  26. 26. Senousy H, Beakes GW, Hack E. Phylogenetic placement of Botryococcus braunii Trebouxiophyceae and Botryococcus sudeticus isolate UTEX 2629 Chlorophyceae. Journal of Phycology. 2004;40:412-423. DOI: 10.1046/j.1529-8817.2004.03173.x
  27. 27. Gouveia D, Lian J, Steinert G, Smidt H, Sipkema D, Wijffels RH, et al. Associated bacteria of Botryococcus braunii Chlorophyta. Peer J. 2019;7:e6610. DOI: 10.7717/peerj.6610
  28. 28. Templar J, Largeau C, Casadevall E. Mechanism of non-isoprenoid hydrocarbon biosynthesis in Botryococcus braunii. Phytochemistry. 1984;23:1017-1028. DOI: 10.1016/S0031-9422(00)82602-3
  29. 29. Templier J, Largeau C, Casadevall E. Effect of various inhibitors on biosynthesis of non-isoprenoid hydrocarbons in Botryococcus braunii. Phytochemistry. 1987;26:377-383. DOI: 10.1016/S0031-9422(00)81418-1
  30. 30. Dennis MW, Kolattukud PE. Alkane biosynthesis by decarbonylation of aldehyde catalyzed by a microsomal preparation from Botryococcus braunii. Archives of Biochemistry and Biophysics. 1991;287:268-275. DOI: 10.1016/0003-9861(91)90478-2
  31. 31. Unkefer CJ, Sayre RT, Magnuson JK, Anderson DB, Baxter I, Blaby IK, et al. Review of the algal biology program within the national alliance for advanced biofuels and bioproducts. Algal Research. 2017;22:187-215. DOI: 10.1016/j.algal.2016.06.002
  32. 32. Sato Y, Ito Y, Okada S, Murakami M, Abe H. Biosynthesis of the triterpenoids, botryococcenes and tetramethylsqualene in the B race of Botryococcus braunii via the non-mevalonate pathway. Tetrahedron Letters. 2003;44:7035-7037. DOI: 10.1016/S0040-4039(03)01784-2
  33. 33. Lohr M, Schwender J, Polle JE. Isoprenoid biosynthesis in eukaryotic phototrophs: A spotlight on algae. Plant Science. 2012;185:9-22. DOI: 10.1016/j.plantsci.2011.07.018
  34. 34. Molnár I, Lopez D, Wisecaver JH, Devarenne TP, Weiss TL, Pellegrini M, et al. Bio-crude transcriptomics: Gene discovery and metabolic network reconstruction for the biosynthesis of the terpenome of the hydrocarbon oil-producing green alga, Botryococcus braunii race B Showa. BMC Genomics. 2012;13:576. DOI: 10.1186/1471-2164-13-576
  35. 35. Matsushima D, Jenke-Kodama H, Sato Y, Fukunaga Y, Sumimoto K, Kuzuyama T, et al. The single cellular green microalga Botryococcus braunii, race B possesses three distinct 1-deoxy-d-xylulose 5-phosphate synthases. Plant Science. 2012;185:309-320. DOI: 10.1016/j.plantsci.2012.01.002
  36. 36. Grung M, Metzger P, Liaaen-Jensen S. Primary and secondary carotenoids in two races of the green alga Botryococcus braunii. Biochemical Systematics and Ecology. 1989;17:263-269. DOI: 10.1016/0305-1978(89)90001-X
  37. 37. Okada S, Matsuda H, Murakami M, Yamaguchi K. Botryoxanthin A, a member of a new class of carotenoids from the green microalga Botryococcus braunii Berkeley. Tetrahedron Letters. 1996;37:1065-1068. DOI: 10.1016/0040-4039(95)02349-6
  38. 38. Okada S, Tonegawa I, Matsuda H, Murakami M, Yamaguchi K. Braunixanthins 1 and 2, new carotenoids from the green microalga Botryococcus braunii. Tetrahedron. 1997;53:11307-11316. DOI: 10.1016/S0040-4020(97)00705-9
  39. 39. Takaichi S. Carotenoids in algae: Distributions, biosynthesis and functions. Marine Drugs. 2011;9:1101-1118. DOI: 10.3390/md9061101
  40. 40. Okada S, Devarenne TP, Chappell J. Molecular characterization of squalene synthase from the green microalga Botryococcus braunii, race B. Archives of Biochemistry and Biophysics. 2000;373:307-317. DOI: 10.1006/abbi.1999.1568
  41. 41. Niehaus TD, Okada S, Devarenne TP, Watt DS, Sviripa V, Chappell J. Identification of unique mechanisms for triterpene biosynthesis in Botryococcus braunii. PNAS. 2011;108:12260-12265. DOI: 10.1073/pnas.1106222108
  42. 42. Okada S, Devarenne TP, Murakami M, Abe H, Chappell J. Characterization of botryococcene synthase enzyme activity, a squalene synthase-like activity from the green microalga Botryococcus braunii, Race B. Archives of Biochemistry and Biophysics. 2004;422:110-118. DOI: 10.1016/j.abb.2003.12.004
  43. 43. Niehaus TD, Kinis S, Okada S, Yeo Y, Bell SA, Cui P, et al. Functional identification of triterpene methyltransferases from Botryococcus braunii Race B. The Journal of Biological Chemistry. 2012;287:8163-8173. DOI: 10.1074/jbc.M111.316059
  44. 44. Tatli M, Naik M, Okada S, Dangott LJ, Devarenne TP. Isolation and characterization of cyclic C33 botryococcenes and a trimethylsqualene isomer from Botryococcus braunii race B. Journal of Natural Products. 2017;80:953-958. DOI: 10.1021/acs.jnatprod.6b00934
  45. 45. Uchida H, Sumimoto K, Ferriols VME, Imou K, Saga K, Furuhashi K, et al. Isolation and characterization of two squalene epoxidase genes from Botryococcus braunii, Race B. PLoS One. 2015;10:e0122649. DOI: 10.1371/journal.pone.0122649
  46. 46. Browne DR, Jenkins J, Schmutz J, Shu S, Barry K, Grimwood J, et al. Draft nuclear genome sequence of the liquid hydrocarbon–accumulating green microalga Botryococcus braunii race B Showa. Genome Announcements. 2017;5:e00215-e00217. DOI: 10.1128/genomeA.00215-17
  47. 47. Brooks J, Shaw G. Sporopollenin: A review of its chemistry, palaeochemistry and geochemistry. Grain. 1978;17:91-97. DOI: 10.1080/00173137809428858
  48. 48. Rozema J, Blokker P, Fuertes MM, Broekman R. UV-B absorbing compounds in present-day and fossil pollen, spores, cuticles, seed coats and wood: Evaluation of a proxy for solar UV radiation. Photochemical & Photobiological Sciences. 2009;8:1233-1243. DOI: 10.1039/B904515E
  49. 49. Jin J, Dupré C, Yoneda K, Watanabe MM, Legrand J, Grizeau D. Characteristics of extracellular hydrocarbon-rich microalga Botryococcus braunii for biofuels production: Recent advances and opportunities. Process Biochemistry. 2016;51:1866-1875. DOI: 10.1016/j.procbio.2015.11.026
  50. 50. Nagano S, Yamamoto S, Nagakubo M, Atsumi K, Watanabe MM. Physical properties of hydrocarbon oils produced by Botryococcus braunii: Density, kinematic viscosity, surface tension, and distillation properties. Procedia Environmental Sciences. 2012;15:73-79. DOI: 10.1016/j.proenv.2012.05.012
  51. 51. Yamamoto S, Mandokoro Y, Nagano S, Nagakubo M, Atsumi K, Watanabe MM. Catalytic conversion of Botryococcus braunii oil to diesel fuel under mild reaction conditions. Journal of Applied Phycology. 2014;26:55-64. DOI: 10.1007/s10811-013-0065-9
  52. 52. Wolf FR, Nemethy EK, Blanding JH, Bassham JA. Biosynthesis of unusual acyclic isoprenoids in the alga Botryococcus braunii. Phytochemistry. 1985;24:733-737. DOI: 10.1016/S0031-9422(00)84886-4
  53. 53. Rao AR, Dayananda C, Sarada R, Shamala TR, Ravishankar GA. Effect of salinity on growth of green alga Botryococcus braunii and its constituents. Bioresource Technology. 2007;98:560-564. DOI: 10.1016/j.biortech.2006.02.007
  54. 54. Yoshimura T, Okada S, Honda M. Culture of the hydrocarbon producing microalga Botryococcus braunii strain Showa: Optimal CO2, salinity, temperature, and irradiance conditions. Bioresource Technology. 2013;133:232-239. DOI: 10.1016/j.biortech.2013.01.095
  55. 55. Gouveia JD, Ruiz J, van den Broek LA, Hesselink T, Peters S, Kleinegris DM, et al. Botryococcus braunii strains compared for biomass productivity, hydrocarbon and carbohydrate content. Journal of Biotechnology. 2017;248:77-86. DOI: 10.1016/j.jbiotec.2017.03.008
  56. 56. Tsutsumi S, Saito Y, Matsushita Y, Aoki H. Investigation of colony disruption for hydrocarbon extraction from Botryococcus braunii. Fuel Processing Technology. 2018;172:36-48. DOI: 10.1016/j.fuproc.2017.12.004
  57. 57. Jackson BA, Bahri PA, Moheimani NR. Response of Botryococcus braunii to repetitive non-destructive extraction of lipids with heptane. In: Chemeca 2018 Conference. 30 September–3 October 2018. Queenstown, NZ: Institution of Chemical Engineers; 2018. pp. 89.1-89.9. Available from: https://search.informit.com.au/documentSummary;dn=047836542036386;res=IELENG
  58. 58. Mehta P, Jackson BA, Nwoba EG, Vadiveloo A, Bahri PA, Mathur AS, et al. Continuous non-destructive hydrocarbon extraction from Botryococcus braunii BOT-22. Algal Research. 2019;41:101537. DOI: 10.1016/j.algal.2019.101537
  59. 59. Chaudry S, Bahr PA, Moheimani NR. Techno-economic analysis of milking of Botryococcus braunii for renewable hydrocarbon production. Algal Research. 2018;31:194-203. DOI: 10.1016/j.algal.2018.02.011
  60. 60. Tasić MB, Pinto LFR, Klein BC, Veljković VB, Filh RM. Botryococcus braunii for biodiesel production. Renewable and Sustainable Energy Reviews. 2016;64:260-270. DOI: 10.1016/j.rser.2016.06.009
  61. 61. Xu L, Liu R, Wang F, Liu CZ. Development of a draft-tube airlift bioreactor for Botryococcus braunii with an optimized inner structure using computational fluid dynamics. Bioresource Technology. 2012;119:300-305. DOI: 10.1016/j.biortech.2012.05.123
  62. 62. Baba M, Kikuta F, Suzuki I, Watanabe MM, Shiraiwa Y. Wavelength specificity of growth, photosynthesis, and hydrocarbon production in the oil-producing green alga Botryococcus braunii. Bioresource Technology. 2012;109:266-270. DOI: 10.1016/j.biortech.2011.05.059
  63. 63. Casadevall E, Dif D, Largeau C, Gudin C, Chaumont D, Desanti O. Studies on batch and continuous cultures of Botryococcus braunii: Hydrocarbon production in relation to physiological state, cell ultrastructure, and phosphate nutrition. Biotechnology and Bioengineering. 1985;27:286-295. DOI: 10.1002/bit.260270312
  64. 64. Ge Y, Liu J, Tian G. Growth characteristics of Botryococcus braunii 765 under high CO2 concentration in photobioreactor. Bioresource Technology. 2011;102:130-134. DOI: 10.1016/j.biortech.2010.06.051
  65. 65. Liu J, Ge Y, Cheng H, Wu L, Tian G. Aerated swine lagoon wastewater: A promising alternative medium for Botryococcus braunii cultivation in open system. Bioresource Technology. 2013;139:190-194. DOI: 10.1016/j.biortech.2013.04.036
  66. 66. Talukdar J, Kalita MC, Goswami BC. Characterization of the biofuel potential of a newly isolated strain of the microalga Botryococcus braunii Kützing from Assam, India. Bioresource Technology. 2013;149:268-275. DOI: 10.1016/j.biortech.2013.09.057
  67. 67. Ashokkumar V, Rengasamy R. Mass culture of Botryococcus braunii Kutz. under open raceway pond for biofuel production. Bioresource Technology. 2012;104:394-399. DOI: 10.1016/j.biortech.2011.10.093
  68. 68. Khatri W, Hendrix R, Niehaus T, Chappell J, Curtis WR. Hydrocarbon production in high density Botryococcus braunii race B continuous culture. Biotechnology and Bioengineering. 2014;111:493-503. DOI: 10.1002/bit.25126
  69. 69. Zhang H, Wang W, Li Y, Yang W, Shen G. Mixotrophic cultivation of Botryococcus braunii. Biomass and Bioenergy. 2011;35:1710-1715. DOI: 10.1016/j.biombioe.2011.01.002
  70. 70. Rao AR, Ravishankar GA, Sarada R. Cultivation of green alga Botryococcus braunii in raceway, circular ponds under outdoor conditions and its growth, hydrocarbon production. Bioresource Technology. 2012;123:528-533. DOI: 10.1016/j.biortech.2012.07.009
  71. 71. Bazaes J, Sepulveda C, Acién FG, Morales J, Gonzales L, Rivas M, et al. Outdoor pilot-scale production of Botryococcus braunii in panel reactors. Journal of Applied Phycology. 2012;24:1353-1360. DOI: 10.1007/s10811-012-9787-3
  72. 72. Ozkan A, Kinney K, Katz L, Berberoglu H. Reduction of water and energy requirement of algae cultivation using an algae biofilm photobioreactor. Bioresource Technology. 2012;114:542-548. DOI: 10.1016/j.biortech.2012.03.055
  73. 73. Cheng P, Ji B, Gao L, Zhang W, Wang J, Liu T. The growth, lipid and hydrocarbon production of Botryococcus braunii with attached cultivation. Bioresource Technology. 2013;138:95-100. DOI: 10.1016/j.biortech.2013.03.150
  74. 74. Ashokkumar V, Agila E, Sivakumar P, Salam Z, Rengasamy R, Ani FN. Optimization and characterization of biodiesel production from microalgae Botryococcus grown at semi-continuous system. Energy Conversion and Management. 2014;88:936-946. DOI: 10.1016/j.enconman.2014.09.019
  75. 75. Nascimento IA, Marques SSI, Cabanelas ITD, Pereira SA, Druzian J, De Souza CO, et al. Screening microalgae strains for biodiesel production: Lipid productivity and estimation of fuel quality based on fatty acids profiles as selective criteria. Bioenergy Research. 2013;6:1-13. DOI: 10.1007/s12155-012-9222-2
  76. 76. Fan JY, Chiu HC, Wu JT, Chiang YR, Hsu SH. Fatty acids in Botryococcus braunii accelerate topical delivery of flurbiprofen into and across skin. International Journal of Pharmaceutics. 2004;276:163-173. DOI: 10.1016/j.ijpharm.2004.02.026
  77. 77. Choi GG, Kim BH, Ahn CY, Oh HM. Effect of nitrogen limitation on oleic acid biosynthesis in Botryococcus braunii. Journal of Applied Phycology. 2011;23:1031-1037. DOI: 10.1007/s10811-010-9636-1
  78. 78. Ruangsomboon S. Effect of light, nutrient, cultivation time and salinity on lipid production of newly isolated strain of the green microalga, Botryococcus braunii KMITL 2. Bioresource Technology. 2012;9:261-265. DOI: 10.1016/j.biortech.2011.07.025
  79. 79. Ruangsomboon S. Effects of different media and nitrogen sources and levels on growth and lipid of green microalga Botryococcus braunii KMITL and its biodiesel properties based on fatty acid composition. Bioresource Technology. 2015;191:377-384. DOI: 10.1016/j.biortech.2015.01.091
  80. 80. Tran HL, Kwon JS, Kim ZH, Oh Y, Lee CG. Statistical optimization of culture media for growth and lipid production of Botryococcus braunii LB572. Biotechnology and Bioprocess Engineering. 2010;15:277-284. DOI: 10.1007/s12257-009-0127-7
  81. 81. Yeesang C, Cheirsilp B. Low-cost production of green microalga Botryococcus braunii biomass with high lipid content through mixotrophic and photoautotrophic cultivation. Applied Biochemistry and Biotechnology. 2014;174:116-129. DOI: 10.1007/s12010-014-1041-9
  82. 82. Cabanelas ITD, Marques SSI, de Souza CO, Druzian J, Nascimento IA. Botryococcus, what to do with it? Effect of nutrient concentration on biorefinery potential. Algal Research. 2015;11:43-49. DOI: 10.1016/j.algal.2015.05.009
  83. 83. Pérez-Mora LS, Matsudo MC, Cezare-Gomes EA, Carvalho J. An investigation into producing Botryococcus braunii in a tubular photobioreactor. Journal of Chemical Technology and Biotechnology. 2016;91:3053-3060. DOI: 10.1002/jctb.4934
  84. 84. Wijihastuti RS, Moheimani NR, Bahri PA, Cosgrove JJ, Watanabe MM. Growth and photosynthetic activity of Botryococcus braunii biofilms. Journal of Applied Phycology. 2017;29:1123-1134. DOI: 10.1007/s10811-016-1032-z
  85. 85. Pangestuti R, Kim SK. Biological activities and health benefit effects of natural pigments derived from marine algae. Journal of Functional Foods. 2011;3:255-266. DOI: 10.1016/j.jff.2011.07.001
  86. 86. Ambati RR, Gogisetty D, Gokare RA, Ravi S, Bikkina PN, Su Y, et al. Botryococcus as an alternative source of carotenoids and its possible applications—An overview. Critical Reviews in Biotechnology. 2018;38:541-558. DOI: 10.1080/07388551.2017.1378997
  87. 87. Mulders KJ, Lamers PP, Martens DE, Wijffels RH. Phototrophic pigment production with microalgae: Biological constraints and opportunities. Journal of Phycology. 2014;50:229-242. DOI: 10.1111/jpy.12173
  88. 88. Tonegawa I, Okada S, Murakami M, Yamaguchi K. Pigment composition of the green microalga Botryococcus braunii Kawaguchi-1. Fisheries Science. 1998;64:305-308. DOI: 10.2331/fishsci.64.305
  89. 89. Matsuura H, Watanabe MM, Kaya K. Echinenone production of a dark red-coloured strain of Botryococcus braunii. Journal of Applied Phycology. 2012;24:973-977. DOI: 10.1007/s10811-011-9719-7
  90. 90. Grung M, Metzger P. Algal carotenoids 53; secondary carotenoids of algae 4; secondary carotenoids in the green alga Botryococcus braunii, race L, new strain. Biochemical Systematics and Ecology. 1994;22:25-29. DOI: 10.1016/0305-1978(94)90111-2
  91. 91. Okada S, Tonegawa I, Matsuda H, Murakami M, Yamaguchi K. Botryoxanthin B and α-botryoxanthin A from the green microalga Botryococcus braunii Kawaguchi-1. Phytochemistry. 1998;47:1111-1115. DOI: 10.1016/S0031-9422(98)80082-4
  92. 92. Nakamura H, Matsunaga S, Kawagishi H, Okada S. Effects of 2-azahypoxanthine on extracellular terpene accumulations by the green microalga Botryococcus braunii, race B. Algal Research. 2016;20:267-275. DOI: 10.1016/j.algal.2016.10.023
  93. 93. Ambati RR, Ravi S, Aswathanarayana RG. Enhancement of carotenoids in green alga-Botryococcus braunii in various autotrophic media under stress conditions. International Journal of Biomedical and Pharmaceutical Sciences. 2010;4:87-92. Available from: https://www.researchgate.net/profile/Dr_Ranga_Rao_Ambati/publication/235352513
  94. 94. Rao AR, Baskaran V, Sarada R, Ravishanka GA. In vivo bioavailability and antioxidant activity of carotenoids from microalgal biomass—A repeated dose study. Food Research International. 2013;54:711-717. DOI: 10.1016/j.foodres.2013.07.067
  95. 95. Rao AR, Sarada R, Baskaran V, Ravishankar GA. Antioxidant activity of Botryococcus braunii extract elucidated in vitro models. Journal of Agricultural and Food Chemistry. 2006;54:4593-4599. DOI: 10.1021/jf060799j
  96. 96. Buono S, Langellotti AL, Martello A, Bimonte M, Tito A, Carola A, et al. Biological activities of dermatological interest by the water extract of the microalga Botryococcus braunii. Archives of Dermatological Research. 2012;304:755-764. DOI: 10.1007/s00403-012-1250-4
  97. 97. Liu L, Pohnert G, Wei D. Extracellular metabolites from industrial microalgae and their biotechnological potential. Marine Drugs. 2016;14:191. DOI: 10.3390/md14100191
  98. 98. Öner ET. Microbial production of extracellular polysaccharides from biomass. In: Fang Z, editor. Pretreatment Techniques for Biofuels and Biorefineries. Berlin Heidelberg: Springer; 2013. pp. 35-56. DOI: 10.1007/978-3-642-32735-3_2
  99. 99. Brockhausen I. The role of galactosyltransferases in cell surface functions and in the immune system. Drug News & Perspectives. 2006;19:401-409. DOI: 10.1358/dnp.2006.19.7.1021491
  100. 100. Matuso N, Kaneko S, Atsushi K, Kobayashi H, Kusakabes I. Purification and characterization and gene cloning of two α-l-arabinofuranosides from Streptomyces chartreusis GS 901. Biochemical Journal. 2000;346:9-15. DOI: 10.1042/bj3460009
  101. 101. Shin HY, Park SY, Sung JH, Kim DH. Purification and characterization of α-l-arabinopyranosidase and α-l-arabinofuranosidase from Bifidobacterium breve K-110, a human intestinal anaerobic bacterium metabolizing ginsenoside Rb2 and Rc. Applied and Environmental Microbiology. 2003;69:7116-7123. DOI: 10.1128/AEM.69.12.7116-7123.2003
  102. 102. Cepák V, Přibyl P. Light intensity and nitrogen effectively control exopolysaccharide production by the green microalga Botryococcus braunii Trebouxiophyceae. Genetics and Plant Physiology. 2018;8:24-37. Available from: http://www.bio21.bas.bg/ippg/bg/wp-content/uploads/2018/08/GPP_8_1-2_2018_24-37.pdf
  103. 103. Dayananda C, Sarada R, Rani MU, Shamala TR, Ravishankar GA. Autotrophic cultivation of Botryococcus braunii for the production of hydrocarbons and exopolysaccharides in various media. Biomass and Bioenergy. 2007;31:87-93. DOI: 10.1016/j.biombioe.2006.05.001
  104. 104. Bayona KCD, Garcés LA. Effect of different media on exopolysaccharide and biomass production by the green microalga Botryococcus braunii. Journal of Applied Phycology. 2014;26:2087-2095. DOI: 10.1007/s10811-014-0242-5
  105. 105. García-Cubero R, Cabanelas ITD, Sijtsma L, Kleinegris DM, Barbosa MJ. Production of exopolysaccharide by Botryococcus braunii CCALA 778 under laboratory simulated Mediterranean climate conditions. Algal Research. 2018;29:330-336. DOI: 10.1016/j.algal.2017.12.003
  106. 106. García-Cubero R, Wang W, Martín J, Bermejo E, Sijtsma L, Togtema A, et al. Milking exopolysaccharides from Botryococcus braunii CCALA778 by membrane filtration. Algal Research. 2018;34:175-181. DOI: 10.1016/j.algal.2018.07.018
  107. 107. Kumar D, Kastanek P, Adhikary SP. Exopolysaccharides from cyanobacteria and microalgae and their commercial application. Current Science. 2018;115:234. Available from: https://www.currentscience.ac.in/Volumes/115/02/0234.pdf
  108. 108. Volkman JK. Acyclic isoprenoid biomarkers and evolution of biosynthetic pathways in green microalgae of the genus Botryococcus. Organic Geochemistry. 2014;75:36-47. DOI: 10.1016/j.orggeochem.2014.06.005
  109. 109. Kavitha G, Kurinjimalar C, Sivakumar K, Kaarthik M, Aravind R, Palani P, et al. Optimization of polyhydroxybutyrate production utilizing waste water as nutrient source by Botryococcus braunii Kütz using response surface methodology. International Journal of Biological Macromolecules. 2016;93:534-542. DOI: 10.1016/j.ijbiomac.2016.09.019
  110. 110. Markou G, Nerantzis E. Microalgae for high-value compounds and biofuels production: A review with focus on cultivation under stress conditions. Biotechnology Advances. 2013;31:1532-1542. DOI: 10.1016/j.biotechadv.2013.07.011
  111. 111. Kavitha G, Kurinjimalar C, Sivakumar K, Palani P, Rengasamy R. Biosynthesis, purification and characterization of polyhydroxybutyrate from Botryococcus braunii K. International Journal of Biological Macromolecules. 2016;89:700-706. DOI: 10.1016/j.ijbiomac.2016.04.086
  112. 112. Arévalo-Gallegos A, Garcia-Perez JS, Carrillo-Nieves D, Ramirez-Mendoza RA, Iqbal HM, Parra-Saldívar R. Botryococcus braunii as a bioreactor for the production of nanoparticles with antimicrobial potentialities. International Journal of Nanomedicine. 2018;13:5591. DOI: 10.2147/IJN.S174205
  113. 113. Sawayama S, Minowa T, Yokoyama SY. Possibility of renewable energy production and CO2 mitigation by thermochemical liquefaction of microalgae. Biomass and Bioenergy. 1999;17:33-39. DOI: 10.1016/S0961-9534(99)00019-7
  114. 114. Ruangsomboon S, Sornchai P, Prachom N. Enhanced hydrocarbon production and improved biodiesel qualities of Botryococcus braunii KMITL 5 by vitamins thiamine, biotin and cobalamin supplementation. Algal Research. 2018;29:159-169. DOI: 10.1016/j.algal.2017.11.028
  115. 115. Calderón NDG, Bayona KCD, Garcés LA. Immobilization of the green microalga Botryococcus braunii in polyester wadding: Effect on biomass, fatty acids, and exopolysaccharide production. Biocatalysis and Agricultural Biotechnology. 2018;14:80-87. DOI: 10.1016/j.bcab.2018.02.006
  116. 116. Al-Hothaly KA. An optimized method for the bio-harvesting of microalgae, Botryococcus braunii, using Aspergillus sp. in large-scale studies. MethodsX. 2018;5:788-794. DOI: 10.1016/j.mex.2018.07.010

Written By

Edmundo Lozoya-Gloria, Xochitl Morales-de la Cruz and Takehiro A. Ozawa-Uyeda

Submitted: 21 June 2019 Reviewed: 24 June 2019 Published: 21 September 2019