Sunteți pe pagina 1din 7

Journal of Biotechnology 162 (2012) 5056

Contents lists available at SciVerse ScienceDirect

Journal of Biotechnology
journal homepage: www.elsevier.com/locate/jbiotec

Cyanobacterial biofuel production


Iara M.P. Machado, Shota Atsumi
Department of Chemistry, University of California, Davis, CA 95616, USA

a r t i c l e

i n f o

a b s t r a c t
The development of new technologies for production of alternative fuel became necessary to circumvent nite petroleum resources, associate rising costs, and environmental concerns due to rising fossil fuel CO2 emissions. Several alternatives have been proposed to develop a sustainable industrial society and reduce greenhouse emissions. The idea of biological conversion of CO2 to fuel and chemicals is receiving increased attention. In particular, the direct conversion of CO2 with solar energy to biofuel by photosynthetic microorganisms such as microalgae and cyanobacteria has several advantages compared to traditional biofuel production from plant biomass. Photosynthetic microorganisms have higher growth rates compared with plants, and the production systems can be based on non-arable land. The advancement of synthetic biology and genetic manipulation has permitted engineering of cyanobacteria to produce non-natural chemicals typically not produced by these organisms in nature. This review addresses recent publications that utilize different approaches involving engineering cyanobacteria for production of high value chemicals including biofuels. Published by Elsevier B.V.

Article history: Received 14 November 2011 Received in revised form 28 January 2012 Accepted 8 March 2012 Available online 16 March 2012 Keywords: Biofuel Cyanobacteria Metabolic engineering Synthetic biology Butanol Fatty acids

1. Introduction Several problems have been generated by the use of fossil fuels for the production of chemicals as well as fuel needed for modern daily life. The consequence of this is an alarming level of pollution that affects the ecosystem (Greenwell et al., 2010). To change our dependency on fossil fuels it is necessary to develop sustainable industrial processes for synthesis of fuels and chemicals from bioresources. Several emerging technologies are being implemented in order to overcome the problem of greenhouse gas (GHG) pollution and replace fossil fuels by promoting viable production of liquid fuels such as fatty acid esters (biodiesel) (Kalscheuer et al., 2006), alkanes (Schirmer et al., 2010), and higher alcohols from renewable sources (Atsumi et al., 2008a,b; Bond-Watts et al., 2011; Shen et al., 2011). However, most of these bioprocesses have a major drawback: they rely on microorganisms that metabolize carbohydrate sources from land-based feedstocks. This competition with land use for food crops results in increased food cost (Scharlemann and Laurance, 2008). Furthermore, biofuel from terrestrial crops (e.g., sugar cane) can cause a great environmental cost compared with fossil fuels. For instance, deforestation to clear land for crops causes emission of large amounts of GHG. Also, nitrogen fertilizers often used to cultivate crops are a known source of nitrous oxide, a greenhouse gas that destroys stratospheric ozone (Scharlemann and Laurance, 2008).

Corresponding author. Tel.: +1 530 752 6595; fax: +1 530 752 8995. E-mail address: atsumi@chem.ucdavis.edu (S. Atsumi). 0168-1656/$ see front matter. Published by Elsevier B.V. doi:10.1016/j.jbiotec.2012.03.005

Plant-derived lignocellulose from agricultural and agroindustrial residues (e.g., wood chips, corn stover, sugarcane, bagasse, rice and wheat straw) is a potentially vast source of renewable energy for fuel production that is not tied to food crops or in limited supply. Lignocellulosic biomass is comprised three main types of carbon-based polymers responsible for plant rigidity and structure: cellulose, hemicellulose and lignin (Madhavan et al., 2011). Cellulose is the major component of lignocellulose and is comprised a linear chain homopolymer of linked glucose units. It exists in crystalline microbrilis that makes its glucose monomers hard to reach (Sanderson, 2011). Hemicelluloses are polymers composed of a range of sugars, whereas lignin is a complex of polymers that are cross-linked and thus difcult to break down. At the moment, the challenge is to nd a viable way to degrade lignocellulosic material into simple sugars available for microorganisms to convert into fuel. There is much research focused on solving this problem, and its solution is a long term goal (Sanderson, 2011). The direct conversion of solar energy into liquid fuel using photosynthetic microorganisms is an attractive alternative to fossil fuels. There are several advantages to using organisms such as microalgae and cyanobacteria: their readily available genetic tools and sequenced genomes; their higher growth rate compared to plants; and their ability to thrive in areas that cannot support agriculture. Utilization of these organisms can provide a way of resolving the potential conict between the use of land for food or for biofuel production. Microalgae provide potential advantages for lipid-derived biofuel, especially biodiesel production, due to their great capacity to convert carbon dioxide into carbon-rich lipids, without competing for arable land necessary for agricultural

I.M.P. Machado, S. Atsumi / Journal of Biotechnology 162 (2012) 5056

51

oleaginous crops. Development in bioprocess engineering and increased knowledge of algal physiology has paved the way for their use in bioenergy applications. Most algal biofuel production work in the U.S. has been carried out by the U.S. Department of Energy (DOE) under auspices of the Aquatic Species Program (ASP) between 1978 and 1996. The result of the effort was summarized in the report published by Sheehan in 1998 (Sheehan et al., 1998). However, despite its great potential, there are many fundamental barriers for industrial production. The feasibility of algal biofuel production depends on lowering cost and increasing efciency in various aspects such as strain improvement, photosynthetic capacity, and harvesting and isolation of products from microalgae (Greenwell et al., 2010; Wijffels and Barbosa, 2010). Cyanobacteria are advantageous organisms for industrial applications as they present fast cell growth, have simple nutrient requirements (mainly water, sunlight and CO2 (Rufng, 2011), are naturally transformable and thus have the potential to be genetically engineered (Golden et al., 1987; Heidorn et al., 2011; Huang et al., 2010; Koksharova and Wolk, 2002; Rufng, 2011). Due to their natural diversity, the capacity of cyanobacteria to grow in a variety of locations, even those unt for agriculture, could be exploited for biofuel production. The possibility of biofuel production at sites near the location of fuel use would also reduce transportation costs. Here, we summarize recent approaches using cyanobacteria as a model organism engineered to convert CO2 to fuels and other products of commercial interest.

2. Isobutyraldehyde production from Synechococcus elongatus PCC 7942 using a keto acid pathway Isobutyraldehyde is used primarily as a chemical intermediate to produce plasticizers, glycols, essential amino acids, avor and fragrance, polymers, insecticides and isobutanol. It is a good target for biosynthetic production by cyanobacteria because its physicochemical properties: low boiling point of 63 C and a vapor pressure of 66 mmHg (at 4.4 C) make it easier to purify during production. The advantage of stripping out isobutyraldehyde is that it also reduces its cytotoxicity, allowing for long-term production. The production of isobutyraldehyde by the cyanobacterium S. elongatus has been reported (Fig. 1) (Atsumi et al., 2009). The 2ketoacid pathway was selected for engineering and expression in a mutant S. elongatus strain to synthesize isobutyraldehyde. 2ketoacids are intermediates in amino acid biosynthesis pathways and can be converted to aldehydes by 2-ketoacid decarboxylase (Fig. 1) (Atsumi et al., 2008b). The ketoacid decarboxylase gene kivd from Lactobacillus lactis was integrated into the S. elongatus genome and expressed under control of the isopropyl-beta-dthiogalactoside (IPTG) inducible Ptrc promoter (Brosius et al., 1985). To direct the ux of carbon toward synthesis of the ketoacid precursor, 2-ketoisovalerate, three genes, alsS (Bacillus subtilis), and ilvCD (Escherichia coli) from the pyruvate-valine biosynthesis pathway were expressed and deemed functional in Synechococcus elongatus PCC7942. The resulting strain was capable of producing 0.723 g/L isobutyraldehyde in 12 days. The enzyme responsible for CO2 xation in the CalvinBensonBassaham cycle is ribulose 1,5bisphosphate carboxylase/oxygenase (Rubisco), encoded by rbcLS (Watson and Tabita, 1997) which is an extremely slow catalyst. Its carboxylation activity directly competes with its oxygenase activity (Andersson and Backlund, 2008) and consequently results in compromised fuel production. To improve CO2 xation, extra copies of rbcLS genes from S. elongatus strain PCC6301 were introduced into the S. elongatus PCC7942-derived production strain, downstream of the endogenous rbcLS genes. The overexpression of rbcLS did not increase the rate of photosynthesis, however Rubisco activity

Fig. 1. The pathway for isobutyraldehyde and isobutanol production in S. elongatus PCC7942. The engineered isobutyraldehyde production pathway consists of four enzymatic steps from pyruvate: AlsS: acetolactate synthase; IlvC: acetohydroxy acid isomeroreductase; IlvD: dihydroxy-acid dehydratase; Kdc: 2-ketoacid decarboxylase. Isobutanol production pathway consists of one enzymatic step from Isobutyraldehyde; Adh: alcohol dehydrogenase.

was around 1.4-fold higher and the isobutyraldehyde production increased to 1.1 g/L over 8 days (Table 1). The productivity of isobutyraldehyde per land area is ve to six times better than estimates for corn and cellulosic ethanol production (Sheehan, 2009). 2.1. Isobutanol production from S. elongatus using a keto acid pathway Isobutanol is a higher alcohol used as a solvent in the manufacture of pesticides, avor and fragrances, and also used to produce corrosion inhibitor additives for lubricant oils. Recently, isobutanol has been designated as a promising gasoline substitute (Atsumi et al., 2008b). Compared to ethanol, isobutanol is a better candidate for gasoline replacement due to its low hygroscopicity, higher energy density and high compatibility with current infrastructure. Isobutanol and 1-butanol share similar physicochemical properties, except that isobutanol has a higher octane rating (RON = 113, MON = 94 for isobutanol compared to RON = 96, MON = 78 for 1-butanol; Wallner et al., 2009). However,

52

I.M.P. Machado, S. Atsumi / Journal of Biotechnology 162 (2012) 5056

Table 1 Chemical products synthesized by engineered cyanobacteria. Product 1-Butanol Fatty acid Isoprene Isobutyraldehyde Isobutanol Titer 13.16 mg L 197 mg L1 50 g g1 DCW day 1.1 g/L 450 mg L1
1

Productivity rate 78.33 g L


1

Host organism S. elongatus PCC7942 Synechocystis sp. PCC6803 Synechocystis sp. PCC6803 S. elongatus PCC7942 S. elongatus PCC7942

Reference Lan and Liao (2011) Liu et al. (2011b) Lindberg et al. (2010) Atsumi et al. (2009) Atsumi et al. (2009)

2.08 g g h 6.23 mg L1 h1 3.12 mg L1 h1

no organism can naturally synthesize isobutanol at high yield and productivity. Recently, a 2-ketoacid-based pathway was developed by Atsumi et al. (2008b) (Fig. 1). 2-Ketoacids are intermediates in amino acid biosynthesis pathways and can be converted to alcohols by 2-ketoacid decarboxylase and alcohol dehydrogenase (Fig. 1) (Atsumi et al., 2008b). An l-valine precursor, 2-ketoisovalerate, was utilized to produce isobutanol. Several synthetic approaches to produce isobutanol from carbohydrate compounds using metabolic engineering have been reported in different organisms including Corynebacterium glutamicum (Smith et al., 2010), Clostridium cellulolyticum (Higashide et al., 2011), and Saccharomyces cerevisiae (Chen et al., 2011). To test whether cyanobacteria are capable of direct synthesis of isobutanol, three alcohol dehydrogenases were tested to convert isobutyraldehyde to isobutanol. YqhD from E. coli (Sulzenbacher et al., 2004) showed the best performance (Atsumi et al., 2009). The strain expressing the l-valine pathway (ALS-IlvCD) and alcohol pathway (Kivd and YqhD) produced 0.450 g/L isobutanol in 6 days (Table 1 and Fig. 1). 2.2. 1-Butanol production from S. elongatus using a modied CoA-dependent pathway Another higher-chain alcohol of interest as a biofuel and chemical feedstock is 1-butanol. This is a primary alcohol produced industrially from the petrochemical feedstock propylene and used as an intermediate in the production of butyl esters. It is also used as a solvent for the extraction of essential oils, a solvent for paints, natural and synthetic resins and gums. 1-Butanol has been proposed as a substitute for diesel fuel and gasoline because of its low hygroscopicity and energy content (27 MJ/L), which is similar to gasoline (32 MJ/L). 1-Butanol can also be produced as a byproduct of microbial fermentation processes. At the beginning of the last century, industrial butanol production was carried out by fermentation of Clostridium acetobutylicum (Jones and Woods, 1986). This process produced mainly acetone, 1-butanol, and ethanol (referred to as the ABE fermentation) at a total concentration of solvents produced ranging from 12 to 20 g/L in batch fermentation starting from 55 to 60 g/L of sugar substrate, it results in solvent yields approximately 0.35 g/g sugar (Green, 2011) with solvent ratio of 6:3:1 (butanolacetoneethanol) (Jones and Woods, 1986; Lee et al., 2008). In Clostridium, 1-butanol is produced by a pathway branched from the butyrate pathway (Jones and Woods, 1986), referred to as the CoA-dependent pathway. Pyruvate, resulting from glycolysis, is cleaved by pyruvate ferredoxin oxidoreductase in the presence of coenzyme A (CoA) to yield carbon dioxide, acetyl-CoA, and reduced ferredoxin. The metabolic pathway from acetyl-CoA to 1-butanol requires the following enzymes: acetyl-CoA acetyltransferase (also called thiolase, THL); -hydroxybutyryl-CoA dehydrogenase (HBD); 3-hydroxybutyrylCoA dehydratase (crotonase, CRT), butyryl-CoA dehydrogenase (BCD); electron transferring protein A and B (ETF-A, ETF-B) and bifunctional butyraldehyde dehydrogenase (BYDH)/butanol dehydrogenase (BDH) (Jones and Woods, 1986; Li et al., 2008). Interest in production of biobutanol from biomass has shone new light on 1-butanol biosynthesis research, redirecting it toward production by non-native and user-friendly hosts that are facultative

anaerobes with fast growth rates, have readily available tools for genetic manipulation, and have well characterized physiological regulation. Butanol pathways have been re-constructed in different non-native hosts including E. coli (Atsumi et al., 2008a; Inui et al., 2008), S. cerevisiae (Steen et al., 2008), Lactobacillus brevis (Berezina et al., 2010), B. subtilis and Pseudomonas putida (Nielsen et al., 2009). The difculty in transferring this pathway to non-native hosts is evident in low 1-butanol titers varying from 0.12 g/L to 1.2 g/L when compared to higher titer production of related compounds such as isobutanol (2050 g/L) (Atsumi et al., 2008b; Baez et al., 2011) and isopropanol (40143 g/L) (Inokuma et al., 2010) using E. coli. However, Shen et al. reported an engineered E. coli that produce 1-butanol at high titer of 15 g/L in ask and 30 g/L in fermentor with product removal in situ, by expressing a modied clostridial 1butanol pathway (Shen et al., 2011). The strategy was to replace the clostridial butyril-CoA dehydrogenase complex (Bcd-EtfAB) by the trans-enoyl-CoA (Ter) from Treponema denticola, because it utilizes only NADH as a direct reducing cofactor and delete host pathways that consume acetyl-CoA (acetate production) and NADH (ethanol, lactate, and succinate production). The resulting NADH and acetylCoA accumulation worked as a driving force to increase substrate for the irreversible Ter reaction toward to 1-butanol production (Shen et al., 2011). Recently it has been reported that production of 1-butanol in cyanobacteria from CO2 was achieved by transferring a modied CoA-dependent pathway into the autotrophic organism (Fig. 2) (Lan and Liao, 2011). Since this CoA-dependent pathway was derived from strict anaerobic bacteria, it was a challenging task to express such a pathway in cyanobacteria, which produce oxygen during photosynthesis. To synthesize 1-butanol, 5 genes were integrated into the S. elongatus PCC7942 genome: hbd, crt, and adhE2 from C. acetobutylicum, ter from T. denticola, and atoB from E. coli. The acetyl-CoA acetyltransferase (AtoB) was selected to replace the acetoacetyl-CoA thiolase (THL) from Clostridium due to its higher specic activity (Duncombe and Frerman, 1976; Shen et al., 2011; Wiesenborn et al., 1988). The activities of these gene products were detected by enzyme assay and Ter and Adh2 activities were lower than that of the other 3 enzymes in the pathway. Ter activity was lower than expected compared to E. coli results (Shen et al., 2011), suggesting that Ter could potentially be a limiting step for 1butanol production in cyanobacteria. The addition of an N-terminal polyhistidine-tag increased Ter activity and butanol production to 13.16 mg L1 (Table 1), suggesting that protein stability or protein folding were issues for Ter in the cyanobacterium host. Clostridium AdhE2 contains an iron-coupled motif and is potentially oxygen sensitive (Fontaine et al., 2002). However, AdhE2 activity was tested in oxic and anoxic conditions, and the resulting comparable activities suggest that low AdhE2 activity may not be related to oxygen sensitivity. The engineered S. elongatus strain did not produce 1-butanol in the presence of oxygen and constant light exposure. To investigate the effect of light and oxygen for 1-butanol production in the cyanobacterium strain, cells were treated with the herbicide 3-(3 ,4 -dichlorophenyl)-1,1-dimethyl urea (DCMU) to inhibit photosynthesis II and thus to generate no oxygen (Metz et al., 1986). The amount of 1-butanol accumulated was proportional to

I.M.P. Machado, S. Atsumi / Journal of Biotechnology 162 (2012) 5056

53

Fig. 2. The pathway for 1-butanol production in S. elongatus PCC7942. The engineered 1-butanol production pathway contains ve enzymatic steps from acetyl-CoA. AtoB: acetyl-CoA acetyltransferase; Hbd: hydroxybutyrl-CoA dehydrogenase; Crt: crotonase; Ter: trans-enoyl-CoA reductase; AdhE2 : bifunctional aldehyde/alcohol dehydrogenase.

DMCU added during light exposure. This suggested that oxygen, rather than light, is an important factor for 1-butanol production in cyanobacteria. 2.3. Isoprene production Isoprene is a 5-carbon, highly volatile hydrocarbon currently used by the rubber industry. It is also a repeating unit in biological substances such as vitamin A and steroid sex hormones (Ruzicka, 1953). Isoprene is articially produced mainly from petroleum as a side product in the production of ethylene (Ruzicka, 1953). Many species of tree naturally produce and emit isoprene (Sharkey et al., 2008). As a small hydrophobic molecule, isoprene is released easily from cell leaves to the atmosphere, but this would be unsuitable for large-scale isolation of isoprene from plants due to difculty in enclosing canopies. However, some microorganisms also naturally produce isoprene (Kuzma et al., 1995) and offer the production advantages of cultivation in enclosed bioreactors, which permits collection and sequestration of the volatile isoprene. Microbial production of isoprene has been recently demonstrated by an engineered Synechocystis sp. PCC 6803 strain (Fig. 3) (Lindberg et al., 2010). Cyanobacteria naturally express the methyl-erythritol-4phosphate (MEP) pathway (Ershov et al., 2002; Okada and Hase, 2005), which is also present in plant plastids (Lichtenthaler, 2000). The MEP pathway in cyanobacteria is important for synthesis of terpenoid-like molecules required for many cell functions including photosynthesis, membrane stability and cellular production of carotenoids (Poliquin et al., 2004). However, the key enzyme that catalyzes the last step of isoprene biosynthesis (IspS) in plants is absent in cyanobacteria. To produce isoprene in Synechocystis, the isoprene synthase gene from kudzu vine (Pueraria montana), an isoprene-emitting plant species (Kesselmeier and Staudt, 1999), was expressed under regulation of the photosystem component psbA2 promoter (Agrawal et al., 2001). The strength of this promoter depends on photon availability. The expression of ispS in Synechocystis was tested under low light (10 mol photons m2 s1 ) and high light (500 mol photons m2 s1 ) growth for 6 h. Transcription of the ispS gene showed a light-intensity-dependent response with accumulation of 50 g isoprene/g dry cell weight/day (Table 1)

(Lindberg et al., 2010). Use of the light-responsive promoter, make it very useful for regulation of heterologous genes in cyanobacteria. 2.4. Fatty acid production in cyanobacteria The gradual exhaustion of fossil-based energy resources is driving chemical industries to search for alternatives to products of the petroleum industry. One of the energy alternatives for petroleumbased diesel fuel is biodiesel. Biodiesel is a clean-burning fuel made by transesterication of triacylglycerides (TAGs) puried from plant oils, yielding fatty acid methyl esters (FAMEs) and fatty acid ethyl esters (FAEEs). Production is dependent on the availability of the specic vegetable oil feedstock and the plant oil must undergo transesterication with short-chain alcohols, usually methanol or ethanol followed by purication. These downstream processes and supply factors can affect the nal price and quality of biodiesel (Kalscheuer et al., 2006). Fatty acids are energy-rich molecules composed of long alkyl chains used by cells for energy storage and chemical production. Fatty acid metabolism in bacteria has been extensively investigated (Cronan and Thomas, 2009). E. coli has been engineered to be an efcient producer of fatty acids (2.5 g/L) by blocking fatty acid degradation, increasing the supplement of malonyl-CoA (an intermediate in fatty acid biosynthesis), and releasing feedback inhibition (Lu et al., 2008). Oil-producing algae are a more prominent biological approach to producing biodiesel because of their capacity to sequester and x CO2 . However, there are still technological barriers and economic issues that hamper the industrial development of high-yield oil-producing algal strains (Robertson et al., 2011; Sheehan et al., 1998). In E. coli, the product of fatty acid biosynthesis is fatty-acyl-ACP (acyl carrier protein), the accumulation of which leads to feedback inhibition of fatty acid biosynthesis. However, high expression of acyl-ACP thioesterase I can deregulate the inhibitory mechanism (Lu et al., 2008). Normally thioesterase I (TesA) is located in the cellular periplasm, but without its signal sequence TesA is located to the cellular cytosol (Cho and Cronan, 1995). E. coli cells producing cytoplasmic TesA have been shown to hydrolize acyl-ACP and overproduce and secrete large amounts of free fatty acids (Cho

54

I.M.P. Machado, S. Atsumi / Journal of Biotechnology 162 (2012) 5056

Fig. 3. The MEP pathway for isoprene production in Synechocystis sp. PCC 6803. The engineered Synechocystis contains endogenous expression of eight enzymes for conversion of pyruvate and glyceraldehyde-3-phosphate (G3P) to dimethylallyl diphosphate (DMAPP) and expression of one heterologous enzyme as the nal step for isoprene production. Dxs: deoxyxylulose 5-phosphate synthase; Dxr: deoxyxylulose 5-phosphate reductoisomerase; IspD: diphosphocytidylyl methylerythritol synthase; IspE: diphosphocytidylyl methylerythritol kinase; IspF: methylerythritol 2,4-cyclodiphosphate synthase; GcpE: hydroxymethylbutenyl diphosphate synthase; IspH: hydroxymethylbutenyl diphosphate reductase; Ipi: isopentenyl diphosphate isomerase; IspS: isoprene synthase.

and Cronan, 1995; Jiang and Cronan, 1994). This concept has been applied to production and secretion of free fatty acids (FFA) in cyanobacteria (Roessler et al., 2009). Thioesterases from bacteria and plants are capable of hydrolyzing acyl-ACP thiol bonds and releasing fatty acids of the desired length (Roessler et al., 2009).

Recently, researchers introduced a codon-optimized acyl carrier protein thioesterase and generated six successive genetic modications in the cyanobacterium Synechocystis sp. PCC6803 (Fig. 4) (Liu et al., 2011b). The mutations resulted in weakened polar cell wall layers due to altered surface proteins and peptidoglycan

Fig. 4. The pathway for fatty acids and alkanes in Synechocystis sp. PCC 6803. Abbreviations used: TE: acyl-acyl carrier protein thioesterase; AAR: acyl-acyl carrier protein reductase; AAD: aldehyde decarboxylase.

I.M.P. Machado, S. Atsumi / Journal of Biotechnology 162 (2012) 5056

55

layers, allowing fast diffusion through phospholipid layers. Removal of cytoplasmic FFA would increase intracellular FFA production by pushing the reaction toward the desired product. The fatty acid secretion yield of the engineered cyanobacterium SD277 was 197 mg L1 of culture (Table 1). A downstream process for more cost-effective and environmentally friendly lipid recovery from cyanobacterial biomass was recently developed. The general process for recovery of algal lipids includes cell disruption, extraction, and separation, and accounts for up to 50% of the total cost of biofuel production. A new proposed process is called the Green Recovery system, which utilizes lipolytic enzymes to degrade membrane lipids into FFAs and lyse the cells (Liu et al., 2011a). Lipase synthesis is under the control of CO2 -limitation-inducible promoters. A strain of Synechocystis sp. engineered for Green Recovery has been reported to release 36.1 1012 mg/cell of fatty acid from degradation of diacylglycerols under CO2 limitation (Liu et al., 2011a). The FFAs produced from CO2 by cyanobacteria can be converted into biodiesel by esterication, and to alkanes and alkenes, through a two-step pathway recently identied in cyanobacteria. Schirmer et al., identied two enzyme families related to alkane biosynthesis, an acyl-acyl-carrier protein reductase and an aldehyde decarbonylase (Schirmer et al., 2010). Both enzymes were coexpressed in E. coli, which allowed for production of alkanes. Their model alkane biosynthesis pathway in cyanobacteria consisted of reduction and decarboxylation, with formation of carbon monoxide as a coproduct. When the model was tested, formate was formed rather than CO during conversion of fatty aldehydes to alkanes (Li et al., 2011; Warui et al., 2011). 3. Conclusions The wakeup call to nd a better replacement for fossil fuel is challenging researchers and companies to look for possibilities other than ethanol. Advances in synthetic biology are making it possible to design organisms that efciently convert solar energy into high-value products. Cyanobacteria have several advantages, most notably their fast growth compared to plants, high photosynthetic efciency, and their capacity for genetic engineering. Reports described above using cyanobacteria as a factory for valuable chemicals contribute to the vision of the viability of industrial photosynthesis. Since the use of cyanobacteria to produce valuable chemicals is still in the initial stage of exploration, there is a long way to go. In addition to product development future research must address strain improvement to achieve high productivity, maintain growth rate and improve cell potential for survival under adverse conditions. These will be important factors to leap from laboratory studies to large-scale and protable biofuel production. Acknowledgments We would like to thank John W.K. Oliver and Christine Rabinovitch-Deere for critical reading of the manuscript. This work was supported by DOE ARPA-E Award (DE-AR0000085). References
Agrawal, G.K., Kato, H., Asayama, M., Shirai, M., 2001. An AU-box motif upstream of the SD sequence of light-dependent psbA transcripts confers mRNA instability in darkness in cyanobacteria. Nucleic Acids Res. 29, 18351843. Andersson, I., Backlund, A., 2008. Structure and function of Rubisco. Plant Physiol. Biochem. 46, 275291. Atsumi, S., Cann, A.F., Connor, M.R., Shen, C.R., Smith, K.M., Brynildsen, M.P., Chou, K.J., Hanai, T., Liao, J.C., 2008a. Metabolic engineering of Escherichia coli for 1butanol production. Metab. Eng. 10, 305311. Atsumi, S., Hanai, T., Liao, J.C., 2008b. Non-fermentative pathways for synthesis of branched-chain higher alcohols as biofuels. Nature 451, 8689.

Atsumi, S., Higashide, W., Liao, J.C., 2009. Direct photosynthetic recycling of carbon dioxide to isobutyraldehyde. Nat. Biotechnol. 27, 11771180. Baez, A., Cho, K.M., Liao, J.C., 2011. High-ux isobutanol production using engineered Escherichia coli: a bioreactor study with in situ product removal. Appl. Microbiol. Biotechnol. 90, 16811690. Berezina, O.V., Zakharova, N.V., Brandt, A., Yarotsky, S.V., Schwarz, W.H., Zverlov, V.V., 2010. Reconstructing the clostridial n-butanol metabolic pathway in Lactobacillus brevis. Appl. Microbiol. Biotechnol. 87, 635646. Bond-Watts, B.B., Bellerose, R.J., Chang, M.C., 2011. Enzyme mechanism as a kinetic control element for designing synthetic biofuel pathways. Nat. Chem. Biol. 7, 222227. Brosius, J., Ere, M., Storella, J., 1985. Spacing of the 10 and 35 regions in the tac promoter. Effect on its in vivo activity. J. Biol. Chem. 260, 35393541. Chen, X., Nielsen, K.F., Borodina, I., Kielland-Brandt, M.C., Karhumaa, K., 2011. Increased isobutanol production in Saccharomyces cerevisiae by overexpression of genes in valine metabolism. Biotechnol. Biofuel 4, 21. Cho, H., Cronan Jr., J.E., 1995. Defective export of a periplasmic enzyme disrupts regulation of fatty acid synthesis. J. Biol. Chem. 270, 42164219. Cronan, J.E., Thomas, J., 2009. Bacterial fatty acid synthesis and its relationships with polyketide synthetic pathways. Methods Enzymol. 459, 395433. Duncombe, G.R., Frerman, F.E., 1976. Molecular and catalytic properties of the acetoacetyl-coenzyme A thiolase of Escherichia coli. Arch. Biochem. Biophys. 176, 159170. Ershov, Y.V., Gantt, R.R., Cunningham Jr., F.X., Gantt, E., 2002. Isoprenoid biosynthesis in Synechocystis sp. strain PCC6803 is stimulated by compounds of the pentose phosphate cycle but not by pyruvate or deoxyxylulose-5-phosphate. J. Bacteriol. 184, 50455051. Fontaine, L., Meynial-Salles, I., Girbal, L., Yang, X., Croux, C., Soucaille, P., 2002. Molecular characterization and transcriptional analysis of adhE2, the gene encoding the NADH-dependent aldehyde/alcohol dehydrogenase responsible for butanol production in alcohologenic cultures of Clostridium acetobutylicum ATCC 824. J. Bacteriol. 184, 821830. Golden, S.S., Brusslan, J., Haselkorn, R., 1987. Genetic engineering of the cyanobacterial chromosome. Methods Enzymol. 153, 215231. Green, E.M., 2011. Fermentative production of butanol the industrial perspective. Curr. Opin. Biotechnol. 22, 337343. Greenwell, H.C., Laurens, L.M., Shields, R.J., Lovitt, R.W., Flynn, K.J., 2010. Placing microalgae on the biofuels priority list: a review of the technological challenges. J. R. Soc. Interface 7, 703726. Heidorn, T., Camsund, D., Huang, H.H., Lindberg, P., Oliveira, P., Stensjo, K., Lindblad, P., 2011. Synthetic biology in cyanobacteria engineering and analyzing novel functions. Methods Enzymol. 497, 539579. Higashide, W., Li, Y., Yang, Y., Liao, J.C., 2011. Metabolic engineering of Clostridium cellulolyticum for production of isobutanol from cellulose. Appl. Environ. Microbiol. 77, 27272733. Huang, H.H., Camsund, D., Lindblad, P., Heidorn, T., 2010. Design and characterization of molecular tools for a synthetic biology approach towards developing cyanobacterial biotechnology. Nucleic Acids Res. 38, 25772593. Inokuma, K., Liao, J.C., Okamoto, M., Hanai, T., 2010. Improvement of isopropanol production by metabolically engineered Escherichia coli using gas stripping. J. Biosci. Bioeng. 110, 696701. Inui, M., Suda, M., Kimura, S., Yasuda, K., Suzuki, H., Toda, H., Yamamoto, S., Okino, S., Suzuki, N., Yukawa, H., 2008. Expression of Clostridium acetobutylicum butanol synthetic genes in Escherichia coli. Appl. Microbiol. Biotechnol. 77, 13051316. Jiang, P., Cronan Jr., J.E., 1994. Inhibition of fatty acid synthesis in Escherichia coli in the absence of phospholipid synthesis and release of inhibition by thioesterase action. J. Bacteriol. 176, 28142821. Jones, D.T., Woods, D.R., 1986. Acetonebutanol fermentation revisited. Microbiol. Rev. 50, 484524. Kalscheuer, R., Stolting, T., Steinbuchel, A., 2006. Microdiesel: Escherichia coli engineered for fuel production. Microbiology 152, 25292536. Kesselmeier, J., Staudt, M., 1999. Biogenic volatile organic compounds (VOC): an overview on emission, physiology and ecology. J. Atmos. Chem. 33, 2388. Koksharova, O.A., Wolk, C.P., 2002. Genetic tools for cyanobacteria. Appl. Microbiol. Biotechnol. 58, 123137. Kuzma, J., Nemecek-Marshall, M., Pollock, W.H., Fall, R., 1995. Bacteria produce the volatile hydrocarbon isoprene. Curr. Microbiol. 30, 97103. Lan, E.I., Liao, J.C., 2011. Metabolic engineering of cyanobacteria for 1-butanol production from carbon dioxide. Metab. Eng. 13, 353363. Lee, S.Y., Park, J.H., Jang, S.H., Nielsen, L.K., Kim, J., Jung, K.S., 2008. Fermentative butanol production by Clostridia. Biotechnol. Bioeng. 101, 209228. Li, F., Hinderberger, J., Seedorf, H., Zhang, J., Buckel, W., Thauer, R.K., 2008. Coupled ferredoxin and crotonyl coenzyme A (CoA) reduction with NADH catalyzed by the butyryl-CoA dehydrogenase/Etf complex from Clostridium kluyveri. J. Bacteriol. 190, 843850. Li, N., Norgaard, H., Warui, D.M., Booker, S.J., Krebs, C., Bollinger Jr., J.M., 2011. Conversion of fatty aldehydes to alka(e)nes and formate by a cyanobacterial aldehyde decarbonylase: cryptic redox by an unusual dimetal oxygenase. J. Am. Chem. Soc. 133, 61586161. Lichtenthaler, H.K., 2000. Non-mevalonate isoprenoid biosynthesis: enzymes, genes and inhibitors. Biochem. Soc. Trans. 28, 785789. Lindberg, P., Park, S., Melis, A., 2010. Engineering a platform for photosynthetic isoprene production in cyanobacteria, using Synechocystis as the model organism. Metab. Eng. 12, 7079.

56

I.M.P. Machado, S. Atsumi / Journal of Biotechnology 162 (2012) 5056 Sharkey, T.D., Wiberley, A.E., Donohue, A.R., 2008. Isoprene emission from plants: why and how. Ann. Bot. 101, 518. Sheehan, J., 2009. Engineering direct conversion of CO(2) to biofuel. Nat. Biotechnol. 27, 11281129. Sheehan, J., Dunahay, T., Benemann, J.R., Roessler, P.G., 1998. A Look Back at the U.S. Department of Energys Aquatic Species Program Biodiesel from Algae. National Renewable Energy Laboratory, Golden, CO. Shen, C.R., Lan, E.I., Dekishima, Y., Baez, A., Cho, K.M., Liao, J.C., 2011. Driving forces enable high-titer anaerobic 1-butanol synthesis in Escherichia coli. Appl. Environ. Microbiol. 77, 29052915. Smith, K.M., Cho, K.M., Liao, J.C., 2010. Engineering Corynebacterium glutamicum for isobutanol production. Appl. Microbiol. Biotechnol. 87, 10451055. Steen, E.J., Chan, R., Prasad, N., Myers, S., Petzold, C.J., Redding, A., Ouellet, M., Keasling, J.D., 2008. Metabolic engineering of Saccharomyces cerevisiae for the production of n-butanol. Microb. Cell Fact. 7, 36. Sulzenbacher, G., Alvarez, K., Van Den Heuvel, R.H., Versluis, C., Spinelli, S., Campanacci, V., Valencia, C., Cambillau, C., Eklund, H., Tegoni, M., 2004. Crystal structure of E. coli alcohol dehydrogenase YqhD: evidence of a covalently modied NADP coenzyme. J. Mol. Biol. 342, 489502. Wallner, T., Miers, S.A., McConnell, S., 2009. A comparison of ethanol and butanol as oxygenates using a direct-injection, spark-ignition engine. J. Eng. Gas Turb. Power 131, 032802. Warui, D.M., Li, N., Norgaard, H., Krebs, C., Bollinger Jr., J.M., Booker, S.J., 2011. Detection of formate, rather than carbon monoxide, as the stoichiometric coproduct in conversion of fatty aldehydes to alkanes by a cyanobacterial aldehyde decarbonylase. J. Am. Chem. Soc. 133, 33163319. Watson, G.M., Tabita, F.R., 1997. Microbial ribulose 1,5-bisphosphate carboxylase/oxygenase: a molecule for phylogenetic and enzymological investigation. FEMS Microbiol. Lett. 146, 1322. Wiesenborn, D.P., Rudolph, F.B., Papoutsakis, E.T., 1988. Thiolase from Clostridium acetobutylicum ATCC 824 and its role in the synthesis of acids and solvents. Appl. Environ. Microbiol. 54, 27172722. Wijffels, R.H., Barbosa, M.J., 2010. An outlook on microalgal biofuels. Science 329, 796799.

Liu, X., Fallon, S., Sheng, J., Curtiss 3rd., R., 2011a. CO2 -limitation-inducible Green Recovery of fatty acids from cyanobacterial biomass. Proc. Natl. Acad. Sci. U. S. A. 108, 69056908. Liu, X., Sheng, J., Curtiss 3rd., R., 2011b. Fatty acid production in genetically modied cyanobacteria. Proc. Natl. Acad. Sci. U. S. A. 108, 68996904. Lu, X., Vora, H., Khosla, C., 2008. Overproduction of free fatty acids in E. coli: implications for biodiesel production. Metab. Eng. 10, 333339. Madhavan, A., Srivastava, A., Kondo, A., Bisaria, V.S., 2011. Bioconversion of lignocellulose-derived sugars to ethanol by engineered Saccharomyces cerevisiae. Crit. Rev. Biotechnol. Metz, J.G., Pakrasi, H.B., Seibert, M., Arntzer, C.J., 1986. Evidence for a dual function of the herbicide-binding D1 protein in photosystem II. FEBS Lett. 205, 269274. Nielsen, D.R., Leonard, E., Yoon, S.H., Tseng, H.C., Yuan, C., Prather, K.L., 2009. Engineering alternative butanol production platforms in heterologous bacteria. Metab. Eng. 11, 262273. Okada, K., Hase, T., 2005. Cyanobacterial non-mevalonate pathway: (E)-4hydroxy-3-methylbut-2-enyl diphosphate synthase interacts with ferredoxin in Thermosynechococcus elongatus BP-1. J. Biol. Chem. 280, 2067220679. Poliquin, K., Ershov, Y.V., Cunningham Jr., F.X., Woreta, T.T., Gantt, R.R., Gantt, E., 2004. Inactivation of sll1556 in Synechocystis strain PCC 6803 impairs isoprenoid biosynthesis from pentose phosphate cycle substrates in vitro. J. Bacteriol. 186, 46854693. Robertson, D.E., Jacobson, S.A., Morgan, F., Berry, D., Church, G.M., Afeyan, N.B., 2011. A new dawn for industrial photosynthesis. Photosynth. Res. 107, 269277. Roessler, P.G., Chen, Y., Liu, B., Dodge, C.N., 2009. Secretion of fatty acids by photosynthetic microorganisms. US Patent 20090298143A1, United States. Rufng, A.M., 2011. Engineered cyanobacteria: teaching an old bug new tricks. Bioeng. Bugs 2, 136149. Ruzicka, L., 1953. The isoprene rule and the biogenesis of terpenic compounds. Experientia 9, 357367. Sanderson, K., 2011. Lignocellulose: chewy problem. Nature 474, S12S14. Scharlemann, J.P., Laurance, W.F., 2008. Environmental science. How green are biofuels? Science 319, 4344. Schirmer, A., Rude, M.A., Li, X., Popova, E., del Cardayre, S.B., 2010. Microbial biosynthesis of alkanes. Science 329, 559562.

S-ar putea să vă placă și