10 1016@j Jbiotec 2019 12 010

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

Journal of Biotechnology 309 (2020) 1–19

Contents lists available at ScienceDirect

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

Review

Enzymatic reactions and pathway engineering for the production of T


renewable hydrocarbons
Juthamas Jaroensuk1, Pattarawan Intasian1, Watsapon Wattanasuepsin,
Nattanon Akeratchatapan, Chatchai Kesornpun, Narongyot Kittipanukul, Pimchai Chaiyen*
School of Biomolecular Science and Engineering, Vidyasirimedhi Institute of Science and Technology (VISTEC), Wangchan Valley, Rayong 21210, Thailand

A R T I C LE I N FO A B S T R A C T

Keywords: Hydrocarbons such as alkanes and alkenes are extensively used as organic compounds for combustion reactions
Biocatalysis and as building block components for the synthesis of numerous materials. Various synthetic enzymatic cascades
Hydrocarbon and engineered metabolic pathways can be used to produce alkanes and alkenes from bio-based materials. An
Enzyme understanding of the native reactions and pathways used by various organisms to synthesize these compounds
Metabolic engineering
together with novel approaches in biocatalysis and synthetic biology have been instrumental in the development
Scale-up
Fatty acid
of methods to produce alkanes and alkenes with reasonable yield. This article discusses the present state of
knowledge regarding hydrocarbon biosynthetic pathways and discusses current mechanistic understanding of
relevant enzymatic reactions in cyanobacteria, aerobic bacteria, insects, algae, and plants. Recent advancements
in metabolic engineering and process scale up for production of hydrocarbons from fatty acids are also discussed.
This technology is important for sustainability, as it provides a clean and eco-friendly method for the future
production of fuels and industrial materials. Further development towards whole cell biocatalysts that are able to
provide good yield with a low production cost may allow countries without big oil reserves to be capable of
producing precursors for the materials industries in the future.

1. Importance of bio-based hydrocarbon production for products such as plastics and detergents. Moreover, these alkenes can
sustainable fuels and materials also be used in biomedical applications i.e. dental polymers and wound
dressings (Thome et al., 2015). The numerous applications using hy-
Hydrocarbons such as alkanes and alkenes are extensively used as drocarbons have made a significant contribution to industrial and
organic compounds for combustion and as building block components urban development since the 19th century (Rogner, 1997)
for the synthesis of various materials. Alkanes are widely used in Currently, most of the hydrocarbons used are derived from fossil
heating and motor fuel applications. The particular usage and appli- fuels. In the US, seventy percent of ethylene monomer production can
cation of a specific hydrocarbon depends on its carbon chain length. be obtained from naphtha steam cracking or ethane thermal cracking
Short-chain alkanes (C2-C4) are gases at atmospheric pressure. They are processes (Ghanta et al., 2014). Fossil fuels are not renewable, and
mainly used in household cooking and for generating electricity by require millions of years to form. Thus, once depleted, they cannot be
incineration. They can be liquified into liquified petroleum gas (LPG) replaced in a short period of time. Renewable sources of hydrocarbons
for use as transportation fuels. Medium-chain alkanes (C5-C17) can be are necessary for global energy security. Moreover, current methods
used as gasoline and diesel fuels for transportation as well as used as used for production of hydrocarbons (petroleum refining of crude oil
solvents for industrial applications. Very long-chain alkanes (> C18) and thermal cracking process) are harsh and unfriendly to the en-
are generally used as lubricating oils. Some of these alkanes are solids vironment. Therefore, an alternative method that is both clean for the
and are thus commonly used as waxes for coating, cosmetic and phar- environment and has economic viability for large-scale production is
maceutical applications. Hydrocarbons also serve as important mono- necessary for future socio-economic sustainability.
mers for polymer production (Fig. 1). Alkenes such as ethene, propene There have been significant advancements in the development of
and 1,3-butadiene, are often used in the manufacture of commodity alternative energy devices during the past decade, challenging the roles


Corresponding author.
E-mail address: pimchai.chaiyen@vistec.ac.th (P. Chaiyen).
1
These authors contributed equally to this work.

https://doi.org/10.1016/j.jbiotec.2019.12.010
Received 3 September 2019; Received in revised form 14 December 2019; Accepted 15 December 2019
Available online 19 December 2019
0168-1656/ © 2019 Elsevier B.V. All rights reserved.
J. Jaroensuk, et al. Journal of Biotechnology 309 (2020) 1–19

Fig. 1. Contribution of bio-based hydrocarbon production as sources of energy and materials. Alkanes and alkenes remain indispensable as power sources (shown in
green) and as chemical building blocks or precursors for materials (shown in orange). Unlike fossil-derived sources which are hazardous for the environment, the bio-
based method offers a clean, renewable and sustainable way for the production of these important chemicals. (For interpretation of the references to colour in this
figure legend, the reader is referred to the web version of this article).

and importance of fossil fuels as energy sources for mobile vehicles. 2. Natural pathways for hydrocarbon production
Currently, the high efficiency of electric vehicle batteries (EVBs) allows
EVBs to serve as power supplies for electric vehicles. In the future, a 2.1. Biosynthesis of hydrocarbons in cyanobacteria
large part of ground transportation is predicted to be diverted from
petroleum-reliant engines to electric engines (Dhameja, 2002; Previous studies revealed that many cyanobacteria can accumulate
Tsiropoulos et al., 2018). However, electric vehicles still require energy hydrocarbon molecules by storing them inside their cells or storing
input from charging stations. While further innovations will likely lead them in membrane lipid form to help cyanobacteria respond to a
to more efficient methods for harnessing power from solar, wind and variety of environmental changes (Wada and Murata, 1998; Wang
hydro sources, the use of hydrocarbons as energy sources for power et al., 2013). Cyanobacteria can convert a variety of carbon feedstocks
stations and for industrial operation will still be indispensable for many such as CO2, glucose or fatty acids to alkane or alkene products using
decades to come (Fig. 1). Thus, the development of methods for hy- their unique metabolic pathways. Two different pathways for alkane or
drocarbon production from renewable sources together with energy alkene biosynthesis exist in cyanobacteria. The first pathway uses acyl-
conversion and improved storage technology will contribute to greater [acyl carrier protein (ACP)] reductase (AAR) to generate fatty aldehyde
sustainability and security of global energy resources (Evans et al., from fatty acyl-ACP and aldehyde-deformylating oxygenase (ADO) to
2009). decarbonylate fatty aldehyde to yield alkanes or alkenes (Fig. 2a). De-
Bio-based methods are alternative, renewable and sustainable ways pending on the substrate captured by AAR, this pathway can yield sa-
for producing alkanes and alkenes. In addition to the benefits of bio- turated alkanes, branched alkanes and alkenes containing internal
based production of hydrocarbons for energy, this technology would double bonds as products. The second pathway utilizes a modular
also be highly beneficial to the production of alkanes and alkenes as multi-domain protein called olefin synthase (OLS) to synthesize alkenes
monomers or precursors for material industries. Therefore, this review from fatty acids. OLS is comprised of three modules (Coates et al.,
article highlights recent findings regarding the enzymatic reactions and 2014b; Mendez-Perez et al., 2011; Zhu et al., 2018). The first module is
metabolic pathways which convert fatty acids to alkanes and alkenes. the activation module which generates fatty acyl-ACP, the second
Here we discuss the natural biosynthetic pathways and enzymatic re- module is an extension module that can extend the fatty acyl-ACP via a
actions found in various organisms and artificial enzymatic cascades malonyl-coenzyme A (malonyl-CoA) unit, and the last module is the
that can be used for hydrocarbon production. The scope of this review alkene formation module which generates the terminal alkene (Fig. 2b)
does not include a well-known and industrial used process of biodiesel (see more discussion in Section 3.3.1).
production as this topic has been covered in great detail in previous
references (Sivaramakrishnan and Incharoensakdi, 2018; Thangaraj
et al., 2018). The report also highlights the recent developments in 2.2. Biosynthesis of hydrocarbons in bacteria
metabolic and process engineering as related to the production of al-
kanes and alkenes by whole cells in large-scale systems. Future per- During the past few years, several hydrocarbon biosynthesis path-
spectives on the technologies required for sustainable production of ways in bacteria have been explored. In bacteria other than cyano-
alkanes and alkenes are also discussed. bacteria, the synthesis of hydrocarbons mainly leads to formation of
alkenes. Two different pathways for alkene biosynthesis exist in these
bacteria. The first pathway used is internal alkene biosynthesis via
head-to-head condensation of fatty acids. In some bacteria such as

2
J. Jaroensuk, et al. Journal of Biotechnology 309 (2020) 1–19

Fig. 2. Overview of enzymatic reaction for hydrocarbon


biosynthesis. (a) Reaction catalyzed by AAR and ADO in
alkane synthesis. AAR, acyl-[acyl carrier protein (ACP)]
reductase; ADO, aldehyde deformylating oxygenase. (b)
Reaction catalyzed by OLS in terminal alkene synthesis.
FAAL, fatty acid-ACP ligase; ACP1, acyl carrier protein 1;
KS, ketosynthase; AT, acyltransferase; KR, ketoreductase;
ACP2, acyl carrier protein 2; ST, sulfotransferase; TE,
thioesterase. (c) Reaction catalyzed by OleABCD in an
internal alkene synthesis via head-to-head condensation.
(d) Reaction catalyzed by fatty acid decarboxylase OleT,
UndA and UndB in terminal alkene synthesis. Pd,
Pseudomonas putida ferredoxin; PNR, Pseudomonas putida
ferredoxin reductase. (e) Reaction catalyzed by acyl-ester
reductase/CYP4Gs and FAR/CYP4Gs in hydrocarbon bio-
synthesis in insects. (f) Photodecarboxylation of fatty acid
by CvFAP. (g) Reaction catalyzed by CER in conversion of
very long-chain fatty acyl-CoA to alkanes and carbon
monoxide in plants.

Micrococcus luteus, Xanthomonas campestris and Shewanella oneidensis, condensation to yield a β-keto acid and the release of 2CoASH. The β-
the OleABCD protein complex catalyzes the formation of internal al- keto acid is subsequently converted into a β-hydroxy acid by OleD
kenes using an initial reaction of OleA to convert β-ketoacyl-CoA (β- through an NADPH-dependent reduction. Formation of the alkene
oxidation intermediate) into an intermediate which is then condensed products is catalyzed by OleC using adenosine triphosphate (ATP)
with another fatty acyl-CoA via non-decarboxylative Claisen (Amelia et al., 2019; Davis et al., 1987; Frias et al., 2010; Sukovich

3
J. Jaroensuk, et al. Journal of Biotechnology 309 (2020) 1–19

et al., 2010) (Fig. 2c). The function of OleB remains unclear. Based on 2009; Harwood and Guschina, 2009; Hu et al., 2008; León-Bañares
sequence homology analysis, OleB belongs to the α/ß hydrolase en- et al., 2004; Liu and Benning, 2013; Malcata, 2011; Sorigué et al., 2016;
zyme family (Frias et al., 2011; Kang and Nielsen, 2017). The second Wijffels et al., 2013). Several algal and microalgal such as Chlamydo-
pathway for alkene biosynthesis is the formation of terminal alkenes. monas, Chlorella, Nannochloropsis, Ostreococcus, and Phaeodactylum
Bacteria from the genus Jeotgalicoccus have been reported to synthesize species were found to be capable of synthesizing long-chain alka(e)nes
1-alkenes through the activity of a fatty acid decarboxylase (OleTJE) by light-dependent pathways (Sorigué et al., 2016).
which is a cytochrome P450 enzyme (CYP450) (Fig. 2d) (see more Despite their valuable ability to produce hydrocarbons, some of the
discussion in Section 3.3.2 and 4.4). Another route for 1-alkene key enzymes responsible for alkane biosynthesis in algae were only just
synthesis is via the reaction of nonheme iron oxidase UndA or mem- identified in 2017. The algal photoenzyme in Chlorella variabilis NC64A
brane-bound desaturase-like enzyme UndB from Pseudomonas. UndA designated as fatty acid photodecarboxylase (CvFAP) uses blue light
and UndB are nonheme mononuclear iron (FeII) oxidases. UndA uses (450 nm) and its cofactor, flavin adenine dinucleotide (FAD) bound at
medium-chain fatty acids (C10-C14) and while UndB uses also short- the enzyme active site, to convert fatty acid substrates into alka(e)nes
chain fatty acids (C6-C16) as substrates. These enzymes convert these and CO2 by decarboxylation (Fig. 2f) (Sorigue et al., 2017). The type of
fatty acid substrates into 1-alkenes through an oxidative decarboxyla- alkane or alkene products from CvFAP usually depends upon the
tion (Fig. 2d) (see more discussion in Section 3.3.3 and 4.4). Both en- properties of the starting substrates (i.e. saturated vs. unsaturated fatty
zymes were first discovered by Prof. Zhang laboratory in UC Berkeley acid). Besides fatty acids, CvFAP can also use short-chain carboxylic
and their sequences show no homology (Rui et al., 2015, 2014). acids to produce short-chain alkanes (Zhang et al., 2019). The enzyme
can convert acetic acid to methane (Zhang et al., 2019). Therefore, this
2.3. Biosynthesis of hydrocarbons in insects reaction is useful and has great versatility in terms of biofuel produc-
tion.
Long-chain hydrocarbons are abundant components found in cuti-
cular lipids located on the surface of all insects. The hydrophobic sur-
face of the insect cuticle plays many important roles in the lives of the 2.5. Biosynthesis of hydrocarbons in plants
insects (Mpuru et al., 1996). The cuticle protects insects from various
conditions such as preventing penetration of inorganic chemicals and The ability of plants to produce very long-chain alkanes is well
acting as a barrier against microorganisms. Insects also use hydro- known. Long-chain alkanes are major compounds in the cuticle located
carbons as chemosensory compounds for chemical communication, and on the primary aerial surface of most terrestrial plants. These com-
most importantly, for regulating the uptake and loss of water (Herman pounds play a key role in minimizing non-stomatal transpiration and
and Zhang, 2016). The cuticular lipids are found widely among various also serve as barriers to protect plants against ultraviolet radiation,
insect species. The cuticles of the majority of insects possess branched pathogens and insects, and provide heat insulation and transpiration
and unsaturated hydrocarbons that all have the same basic structures, control (Bourdenx et al., 2011). Most of the cuticle in plants is located
generally consisting of 19–35 carbon atoms with different ratios of in- on the outer side of epidermal cell walls and consists of a cutin matrix
dividual hydrocarbons in the cuticular lipids varying from 3 % to 95 % as well as cuticular waxes which are abundant in very long-chain fatty
among different species (Lockey, 1988). acids, alcohols, aldehydes, alkanes, ketones, and esters, as well as small
Based on the experiments from Reed et al. (1994) which used mi- amounts of triterpenoids, flavonoids, and sterols (Samuels et al., 2008).
crosomal preparations from the house fly, Musca domestica, fatty acyl- Types and amounts of cuticular wax compounds vary widely among
CoAs can be reduced to aldehydes by the acyl-ester reductase and then plant compartments and species. Their contents in several genus such as
converted to hydrocarbons by the enzyme aldehyde decarbonylase. The Cactaceae, Cupressaceae or Poaceae have been identified (Bi et al., 2005;
carbonyl carbon from the aldehyde is released as CO2 in the microsome Bugalho et al., 2009; Buschhaus et al., 2008; Busta et al., 2016; Dodd
during hydrocarbon formation. The data also demonstrate that the et al., 1998; Dominguez et al., 2011; Dragota and Riederer, 2007;
conversion of aldehyde into hydrocarbon in house fly microsomes re- Diefendorf et al., 2011; Gao and Huang, 2013; Guhling et al., 2005;
quires NADPH and O2 for the activity of the enzyme aldehyde dec- Hegebarth et al., 2017; Hegebarth and Jetter, 2017; Jetter and
arbonylase. The data indicate that the enzyme mechanism may be si- Riederer, 2011; Bird and Gray, 2003; Mihailova et al., 2015; Nadiminti
milar to those of other CYP450 type reactions (Reed et al., 1994). Qiu et al., 2015; Titah et al., 2013; Zhang et al., 2004; Zheng et al., 2005).
et al. (2012) identified that the insect aldehyde decarbonylase is a Interestingly, the data suggest that very long cuticular waxes con-
membrane-bound CYP450 type enzyme. The results from in vitro ana- stituents are limited to only certain hydrocarbons mostly alkanes. Other
lyses using purified microsomes expressing the recombinant fusion compounds such as alcohols or aldehydes chain lengths between C24
protein of CYP450 reductase (CPR) with cytochrome P450 4G1 and C34 tend to be found in cuticular waxes (Bugalho et al., 2009).
(CYP4G1) from Drosophila melanogaster and cytochrome P450 4G2 Although their involvement in cuticle formation is known, the plant
(CYP4G2) from house fly showed that the insect CYP4Gs function as alkane biosynthesis pathway is still not thoroughly understood. The
oxidative decarbonylases which require O2 as a substrate to generate biosynthesis pathway of Arabidopsis thaliana (A. thaliana) is the most
alka(e)ne from aldehydes (Fig. 2e). investigated pathway regarding this aspect. This plant can produce very
Recently, MacLean et al. (2018) reported another pathway in which long-chain (VLC) alkanes dominated by nonacosane (C29 alkane),
insect cuticular hydrocarbons are formed by fatty acyl-CoA reductase which accounts for up to 70 % of the total cuticular wax content in the
(FAR), which converts fatty acyl-CoAs to alcohols. The alcohol moieties leaf (Jetter and Kunst, 2008). A. thaliana was used as a model organism
are then oxidized to aldehydes and to alka(e)nes by CYP4Gs dec- for investigating the mechanisms of alkane biosynthesis in plants. It is
arbonylases (Fig. 2e). As the oxidation of primary alcohols can be currently believed that long-chain fatty acyl-CoAs are modified by the
catalyzed by many types of cytochrome P450 s (MacLean et al., 2018), enzymes in the alkane formation pathway. CER3 reduces long-chain
the ability of CYP4Gs to oxidize primary alcohols to aldehydes and then fatty acyl-CoAs into long-chain fatty aldehydes with predominantly
decarbonylate the aldehydes to alkanes can be expected. even carbon numbers and CER1 enzymes consecutively decarbonylates
the aldehyde to generate an alkane production carbon shorter than the
2.4. Biosynthesis of hydrocarbons in algae substrate (Fig. 2g) (Hegebarth and Jetter, 2017).

Algae and microalgae have been proposed as potential organisms


for use in the synthesis of hydrocarbons from fatty acids because of
their ability to produce plenty of biomass and fatty acids (Beer et al.,

4
J. Jaroensuk, et al. Journal of Biotechnology 309 (2020) 1–19

Fig. 3. Reaction mechanism of CAR catalysis of the reduction of a carboxylic acid to its corresponding aldehyde (Finnigan et al., 2017; Gahloth et al., 2017; Horvat
et al., 2019; Kramer et al., 2018; Qu et al., 2019, 2018; Stolterfoht et al., 2018; Winkler, 2018). Reaction steps consist of activation of the carboxylic acid through
adenylation, thioesterification of the acyl group, and transfer of the acyl thioester from the A-domain to the R-domain.

3. Enzymes used in biocatalysis for production of hydrocarbons site of the A-domain, thus promoting the thioesterification (Finnigan
from fatty acids et al., 2017; Gahloth et al., 2017; Horvat et al., 2019; Kramer et al.,
2018; Qu et al., 2019, 2018; Stolterfoht et al., 2018; Winkler, 2018).
3.1. Enzymes involved in the reduction of fatty acid or fatty acid derivatives The crystal structure of SrCAR during the thioesterification step is
to fatty aldehydes shown in Fig. 4b (Gahloth et al., 2017). The third step is the transfer of
the acyl thioester from the T-domain to the R-domain. This reaction
3.1.1. Carboxylic acid reductase (CAR) step is executed by the swing of the phosphopantetheine arm on the T-
Carboxylic acid reductase (CAR) is an important enzyme that has domain. The final step is the reduction of the acyl thioester inter-
been employed in several synthetic pathways for the production of al- mediate by NADPH and the release of aldehyde and NADP+ products. It
dehydes, key intermediates for the production of hydrocarbons and has been speculated that the aspartate D998 in SrCAR located near the
other fine chemicals (Finnigan et al., 2017; Kramer et al., 2018; Qu NADPH binding site plays a crucial role in regulating the reduction of
et al., 2018; Schwendenwein et al., 2016, 2019; Winkler, 2018; Wood carboxylic acid to aldehyde as a final product. The D998 G variant of
et al., 2017). CAR is a large (MW ∼ 130 kDa) multidomain protein SrCAR showed that the rate of benzoic acid reduction to form benzyl
consisting of an N-terminal adenylation domain (A-domain), a thiola- alcohol was lowered than the wildtype enzyme (Gahloth et al., 2017).
tion domain (T-domain), and a C-terminal reductase domain (R-do- Recently, Stolterfoht et al. (2018) identified a set of amino acid
main) (Finnigan et al., 2017; Kramer et al., 2018; Qu et al., 2018; residues that are important for CAR activity. Mutagenesis of two re-
Stolterfoht et al., 2017, 2018; Winkler, 2018). sidues in the A-domain (H237A and G433A in Neurospora crassa CAR
The mechanism of CAR catalysis of the reduction of a carboxylic (NcCAR)), one residue in the T-domain (S595A in NcCAR, equivalent to
acid to its corresponding aldehyde has been proposed to contain 4 re- S702 in SrCAR), or two residues in the R-domain (Y844A and K848A in
action steps (Fig. 3) (Finnigan et al., 2017; Gahloth et al., 2017; Horvat NcCAR) abolished CAR activity, indicating these residues are essential
et al., 2019; Kramer et al., 2018; Qu et al., 2019, 2018; Stolterfoht et al., for carboxylic acid reduction (Stolterfoht et al., 2018). Moreover, mu-
2018; Winkler, 2018). The first step is the activation of the carboxylic tation of P234A, P285A and E441A in NcCAR increased the specific
acid through adenylation. At this step, the carboxyl group of the acid activity of NcCAR towards pentanoic acid, hexanoic acid, heptanoic
reacts with the α-phosphate of ATP to form an acyl-AMP intermediate acid, octanoic acid substrates, as compared to the wild type NcCAR,
with release of pyrophosphate (PPi). The X-ray crystal structures of suggesting these residues play a key role in substrate binding and
CARs in complex with AMP reveal that the universal conserved lysine substrate preference (Stolterfoht et al., 2018).
residue (K629 in Segniliparus rugosus CAR (SrCAR)), located in the CAR is widely distributed in numerous fungi and actinobacteria.
mobile region of the A-domain (residues 527–654 in SrCAR), promotes Purified CARs from these organisms have been reported to be active
adenylation in the catalytic region of the A-domain by stabilizing and against a broad range of substrates such as C4-C18 fatty acids, hydro-
bringing together ATP and carboxylic acid to facilitate the reaction xyacids, dicarboxylic acids, and (hetero)aromatic acids (Finnigan et al.,
(Finnigan et al., 2017; Gahloth et al., 2017; Horvat et al., 2019; Kramer 2017; Horvat et al., 2019; Kramer et al., 2018; Schwendenwein et al.,
et al., 2018; Qu et al., 2019, 2018; Stolterfoht et al., 2018; Winkler, 2016; Stolterfoht et al., 2017). To date, all of the reported CARs require
2018). The crystal structure of SrCAR during the adenylation step and NADPH and ATP as co-substrates to complete the reduction of car-
the interaction between the AMP-bound domain and the mobile region boxylic acid to aldehyde (Finnigan et al., 2017; Horvat et al., 2019;
of the A-domain is shown in Fig. 4a (Gahloth et al., 2017). The second Kramer et al., 2018; Schwendenwein et al., 2016; Stolterfoht et al.,
step is thioesterification, in which the thiol group of the phospho- 2017). The Km values for NADPH and ATP have been reported in a
pantetheine arm attached to the conserved serine residue (S702 in range of 24–36 μM and 64–84 μM, respectively (Finnigan et al., 2017).
SrCAR) on the T-domain forms a thioester with an acyl group of the Biochemical characterizations of purified CARs from Mycobacterium
acyl-AMP intermediate with concomitant release of AMP. The phos- phlei (MpCAR), Mycobacterium smegmatis (MsCAR), Nocardia iowensis
phopantetheine arm on the T-domain is post-translationally attached by (NiCAR), Nocardia otitidiscaviarum (NoCAR), and Tsukamurella paur-
a phosphopantetheinyl transferase (PPT). In the thioesterification state, ometabola (TpCAR) showed that CARs from these organisms prefer to
structural transformation of the mobile region of the A-domain and the use C8-C12 fatty acids as substrates more than benzoic acid and its
T-domain is required to orient the T-domain to be close to the catalytic derivatives as judged by the kcat/Km value (Finnigan et al., 2017). When

5
J. Jaroensuk, et al. Journal of Biotechnology 309 (2020) 1–19

Fig. 4. (a) The crystal structure of SrCAR during the adenylation step (PDB ID code 5MSW) and (b) the thioesterification step (PDB ID code 5MSS) (Gahloth et al.,
2017). During the adenylation step, the T-domain is located away from the A-domain and the conserved lysine (K629 in SrCAR) rotates into the catalytic site. K629
makes a contact with ATP and catalyzes the adenylation by bringing together ATP and carboxylic acid to achieve formation of the acyl-AMP intermediate. After the
adenylation step, the mobile regions of the A-domain and the T-domain rotate into the form that enables the thioesterification (K629 moves away from the catalytic
site of the A-domain, and S702 moves to be closer to the phosphate of the bound AMP). The residues S702, G532, K528, and, G532 exhibit large movements when the
enzyme switches from the adenylation step to the thioesterification step.

the reactions of benzoic acid substrates were compared, the data steric hindrance of ortho- substituents is thought to be a major cause of
showed that substituted benzoic acids with electron-withdrawing the activity lowering effect in CARs (Finnigan et al., 2017). For benzoic
groups at the para- and meta-positions had lower values of Km/kcat acids with electron-donating groups at the para- and meta-position,
compared to that of benzoic acid without any substituents while ben- activities of CARs were only found to be slightly decreased (Finnigan
zoic acids with ortho-substituents could not be used as substrates for et al., 2017).
CAR. For benzoic acids with electron-donating groups at the ortho-po- Several hydroxyacids (C2-C6) and dicarboxylic acids (C2-C6) have
sition, the absence of activity was observed for MpCAR, MsCAR, NiCAR, also been shown to be substrates for CARs, in which the reactions of
and NoCAR, while a decrease in activity was observed for TpCARs. The hydroxyacids are more favored than those of dicarboxylic acids

6
J. Jaroensuk, et al. Journal of Biotechnology 309 (2020) 1–19

(Kramer et al., 2018). Steady-state kinetic analyses using the purified pentadecane (C15 alkane) as a major product (glucose was used as sole
CARs from Mycobacterium avium (MavCAR), Mycobacterium ar- carbon source), indicating that PmAAR and 7336AAR have higher ac-
omaticivorans (MarCAR), and Kutzneria albida (KaCAR) showed that the tivity toward C16-ACP than other substrates. On the contrary, hepta-
catalytic efficiency of these purified CARs with hydroxyacids and di- decane (C17 alkane) is a primary product produced in the E. coli co-
carboxylic acids increases with longer chain length, but lower than that expressing NpADO and AARs derived from freshwater cyanobacteria
with benzoic acid (Kramer et al., 2018). Overall, the results indicate such as SeAAR, Microcystis aeruginosa (MaAAR), or Thermo-
that CARs prefer electron-rich acid substrates in which a carboxyl group synechococcus elongatus BP-1 (BP-1 AAR). The data imply that C18-ACP
is the only charged group (Finnigan et al., 2017; Kramer et al., 2018). may be the most preferred substrate for AAR derived from fresh water
Additionally, CARs have a higher tolerance towards para- and meta- marine bacteria. However, the type and amount of alkanes produced in
substituents than to ortho-substituents (Finnigan et al., 2017). Recently, hosts cannot be used as absolute indication about substrate preference
Schwendenwein et al. (2019) reported identification of a hot spot re- because the data also depend on the type and amount of substrates that
sidue at the active site of NiCAR which increased its activity towards can be produced by hosts.
meta-and ortho-substituents. Mutagenesis of Q283 to P283 increases
NiCAR affinity to 2‐methoxybenzoic acid, about 9-fold compared to 3.1.3. Fatty acyl-CoA reductase (ACR)
wild type NiCAR. In addition, the catalytic performance of Q283 P Fatty acyl-CoA reductase (ACR) is a key enzyme involved in wax
NiCAR appeared to improve about 3-fold and 4-fold towards 2‐sub- ester biosynthesis in Acinetobacter species (Reiser and Somerville,
stituted (2‐Cl, 2‐Br), and 3‐ and 4‐substituted benzoic acid (3-OMe and 1997). This enzyme can be utilized for generating aldehyde inter-
4-OMe), respectively. mediates in the same manner as CAR. Although ACR can be expressed
NADP+, AMP, and PPi have been shown to inhibit the activity of well in E. coli, it could not be purified to homogeneity. Therefore, no
CARs in vitro. Finnigan et al. (Finnigan et al., 2017) demonstrated that ACR crystal structure is available to date.
NADP+ and AMP inhibit CAR by competing against NADPH (Ki = An enzyme activity assay using ACR in crude lysate from
143 ± 8 μM for NADP+) and ATP (Ki = 8200 ± 900 μM for AMP). Acinetobacter calcoaceticus revealed that ACR catalyzes the formation of
While PPi exhibits mixed type inhibition against ATP (Ki = 220 ± 50 fatty aldehyde by using NADPH for reduction of fatty acyl-CoA (Reiser
μM), it showed competitive inhibition against the 4-methylbenzoic acid and Somerville, 1997). Substrate specificity assays indicated that ACR
substrate (Ki = 340 ± 40 μM). can utilize a broad range of fatty acyl-CoA as substrates (C14 to C22).
However, ACR has the highest activity towards C16:0-CoA, followed by
3.1.2. Acyl-[acyl carrier protein (ACP)] reductase (AAR) C18:0-CoA, C14:0-CoA, C20:0-CoA, and C22:0-CoA respectively. There
Acyl-[acyl carrier protein (ACP)] reductase (AAR) is a key enzyme was no activity observed when the enzyme was tested with substrates of
for alkane biosynthesis in cyanobacteria. It catalyzes NAD(P)H-depen- acyl chain lengths shorter than C12 (Reiser and Somerville, 1997). ACR
dent reduction by reducing fatty acyl-ACP/CoA substrate to its corre- can also be overexpressed in conjunction with ADO to produce alkanes
sponding fatty aldehyde and releasing a free ACP or CoASH (Fig. 2a). in engineered metabolic pathways (Jaroensuk et al., 2019), see more
To date, the AAR crystal structure has not yet been reported. discussion in Section 4.3.
Steady-state kinetic analysis of the purified AAR from Synechococcus
elongatus PCC7942 (SeAAR) revealed that SeAAR catalyzes the reduc- 3.2. Enzymes involved in decarbonylation of fatty aldehyde
tion of fatty acyl-ACP substrate to aldehyde via formation of an acyl-
enzyme thioester intermediate with a cysteine catalytic residue of 3.2.1. Aldehyde deformylating oxygenase (ADO)
SeAAR before the intermediate is reduced to form aldehyde by NADPH. Aldehyde deformylating oxygenase (ADO) is a non-heme di-iron
The SeAAR enzyme is exclusively specific for NADPH enzyme that catalyzes the last step in the biosynthesis of long-chain
(Km = 35.6 ± 4.9 μM) because NADH cannot be used to generate the hydrocarbons. It catalyzes the decarbonylation of fatty aldehydes to
aldehyde product in vitro (Lin et al., 2013). Based on the investigations form Cn-1 alka(e)nes through a redox oxygenation process which re-
of AAR from Nostoc punctiforme PCC73102 (NpAAR) by Warui et al. quires O2 for activity to provide one oxygen atom which is incorporated
(2015), the intermediate was reported to form through the thiol group into formate and also requires a reducing system consisting of NADPH,
of C294 in the active site of NpAAR. ferredoxin (Fd), and ferredoxin reductase (FNR) to provide four elec-
Substrate-specificity analysis demonstrated that AARs can utilize trons per turnover (Buer et al., 2014; Das et al., 2014; Li et al., 2012,
C12-C20 fatty acyl-CoA and C15-C18 fatty acyl-ACP as substrates (Lin 2011; Paul et al., 2013; Rajakovich et al., 2015; Zhang et al., 2013).
et al., 2013; Schirmer et al., 2010; Warui et al., 2015). When tested with The mechanisms of ADO underlying catalysis of decarbonylation of
C8-C20 fatty acyl-CoA, the purified SeAAR showed the highest activities fatty aldehydes to Cn-1 alka(e)nes have been proposed (Fig. 5) (Buer
towards C18:0-CoA (kcat = 0.36 ± 0.023 min−1, and et al., 2014; Das et al., 2014; Li et al., 2012, 2011; Paul et al., 2013;
Km = 31.9 ± 4.2 μM), followed by C20:0-CoA, and C18:1-CoA/C16:0- Rajakovich et al., 2015; Zhang et al., 2013). The reaction of ADO begins
CoA/C12:0-CoA. There was no enzyme activity detected with C8:0-CoA with (I) the aldehyde substrate binds to the di-iron center (Fe2II/II) and
(Lin et al., 2013). Similar substrate preference with respect to acyl forms the intermediate 1 (Fe2II/II-aldehyde). Based on investigations of
chain length was observed when the purified NpAAR was tested with substrate binding to ADO using NMR, Buer et al. (2014) showed that
C18-C12 fatty acyl-ACP. NpAAR exhibits a maximum activity towards the non-hydrated aldehyde is in the aldehyde form when bound to the
C18:0 fatty acyl-ACP (kcat = 0.17 min−1, and Km = 24 ± 6 μM), fol- enzyme. In addition, the X-ray crystal structure of metal bound ADO by
lowed by C16:0-ACP, C15:0-ACP and C18:1-CoA. There was no activity Buer et al. (2014) also revealed that the aldehyde oxygen is ligated to
observed when substrates with acyl chain lengths shorter than C14 Fe2 of the di-iron center, while the water molecule is ligated to Fe1. (II)
were used (Warui et al., 2015). It is worth mentioning that a mono- Molecular O2 binds to the intermediate 1 at the di-iron center (Fe2II/II)
valent cation (K+) and divalent cations (Mg2+, Mn2+, and Ca2+) are and generates the di-iron peroxo species (Fe2III/III) which then nucleo-
required for enhancing the in vitro activity and stability of the purified philically attacks the carbonyl group of the aldehyde. (III-IV) The OeO
SeAARs and NpAAR (Lin et al., 2013; Warui et al., 2015). bond of the intermediate breaks and receives one electron to generate a
Kudo et al. (2016) revealed that AARs derived from marine cya- hemiacetal radical, followed by scission of the C1-C2 bond to yield an
nobacteria and from freshwater cyanobacteria have different substrate alkyl radical. (V) The alkyl radical abstracts one proton and one elec-
specificities. Co-overexpression of AARs derived from marine cyano- tron from the iron-bound water molecule, resulting in formation of an
bacteria such as Prochlorococcus marinus MIT9313 (PmAAR) or Sy- alkane and regeneration of the Fe1IV/Fe2III cluster (Buer et al., 2014;
nechococcus sp. PCC7336 (7336AAR) with Nostoc punctiforme Waugh and Marsh, 2014). Buer et al. (2014) reported that the water
PCC73102 ADO (NpADO) in E. coli resulted in formation of molecule enters the enzyme (PmADO) through the channel formed

7
J. Jaroensuk, et al. Journal of Biotechnology 309 (2020) 1–19

Fig. 5. The mechanism proposed for the decarbonylation of aldehydes to form Cn-1 alka(e)nes catalyzed by ADO (Buer et al., 2014; Das et al., 2014; Li et al., 2012,
2011; Paul et al., 2013; Rajakovich et al., 2015; Zhang et al., 2013). The left and right iron atoms are referred to as Fe1 and Fe2, respectively.

between helices 1 and 3. (VI) The reaction center receives one proton that ADO has broad substrate specificity with respect to acyl chain
and one electron from the reducing system and the alkane product is length. Nevertheless, redesigning the enzyme by site-directed muta-
released. (VII) Formate is released as a final product while the enzyme genesis of the amino acids involved in substrate channeling and the
accepts two electrons from the reducing system to convert di-ferric iron specificity of ADO resulted in the enzyme with significantly improved
(Fe2III/III) to be the active di-ferrous iron (Fe2II/II) center for the sub- catalytic properties (Bao et al., 2016; Khara et al., 2013; Wang et al.,
sequent turnover. The key reductant interacting with ADO during the 2017). For example, the catalytic efficiency (kcat/Km) of the variant
catalytic turnover is reduced Fd which is generated by FNR using A121 F of SeADO towards the C6 aldehyde was improved 3.8-fold
NADPH as the ultimate reductant. Recently, Jaroensuk et al. showed compared to the wild-type enzyme (Bao et al., 2016).
that NADH can also be used by FNR as a reductant to generate reduced
Fd, which in turn can be used to support the activity of ADO. In vitro
analysis using a C14 aldehyde as a substrate indicates that the 3.3. Enzymes involved in the decarboxylation of fatty acids
Fd–NADH–FNR–ADO reconstituted system can generate tridecane as
effectively as the Fd–NADPH–FNR–ADO system (Jaroensuk et al., 3.3.1. Olefin synthase (OLS)
2019). Olefin synthase (OLS) is a modular multi-domain enzyme consisting
To identify residues important for the activity of ADO, site directed of three modules (Fig. 2b) (Chatterjee et al., 2018; Coates et al., 2014a;
mutagenesis of residues identified by X-ray crystal structures were Gu et al., 2009; Mendez-Perez et al., 2011; Moss et al., 2016; Zhu et al.,
carried out (Bao et al., 2016; Jia et al., 2015; Khara et al., 2013; Wang 2018). The first module is the substrate (fatty acid) activation module
et al., 2017). Bao et al. (2016) demonstrated that mutation of residues located at the N-terminus of OLS, comprised of a fatty acyl-ACP ligase
which are located close to the hydrophobic tail of the aldehyde sub- (FAAL) and ACP1. FAAL uses ATP to activate fatty acids to yield fatty
strate (Fig. 6) resulted in variants (Y21R and C70 F in SeADO) that have acyl-AMP, which is then attacked by the SH group of the phospho-
lower activities with medium- and long-chain aldehydes than the wild- pantetheine moiety on ACP1 to generate a fatty acyl-ACP1 intermediate
type enzyme: (I) Y21R exhibits low activities with C10:0 and C12:0 (Fig. 2b, Activation module). This intermediate is later transferred to
aldehydes; (II) Y21R and C70 F exhibit low activities with C14:0-C18:0 the extension module. On this module, the acyl chain of the substrate is
aldehydes; (III) Y21R and C70 F exhibit comparable activities as the elongated by the activities of ketosynthase (KS), acyltransferase (AT),
wild-type enzyme when using short-chain aldehydes as substrates. The ketoreductase (KR), and ACP2. The KS and AT add two carbons from
data clearly indicate that Y21 and C70 are important for medium- to malonyl-CoA to fatty acyl-ACP2 and the β-carbonyl group is reduced to
long-chain substrate selectivity. Wang et al. (2017) reported that mu- form a hydroxy group by the KR domain. The resulting fatty acyl-ACP2
tation of residues close to the di-iron center and active site of the en- is converted into a terminal alkene by the activities of enzymes in the
zyme (Y122F and R62A SeADO, Fig. 6) resulted in a significant decrease terminal alkene formation module, which are sulfotransferase (ST) and
in SeADO activity, with the apparent kcat dropping by 70 % for Y122 F thioesterase (TE). In this step, the ST domain adds a sulfonate group to
and 92 % for R62A. This may due to the disruption of the hydrogen the β-hydroxy group via the sulfonate donor, phosphoadenosinepho-
bonding network around the di-iron cofactor, impairing protein stabi- sphosulfate (PAPS), followed by decarboxylation and hydrolysis of the
lity and activity. In addition, mutation of residues at the protein surface β-sulfate to form the terminal alkene as catalyzed by the TE domain
such as W178R in SeADO Fig. 6 caused the apparent kcat of W178R (Chatterjee et al., 2018; Coates et al., 2014a; Gu et al., 2009; Mendez-
SeADO to increase 3-fold compared to that of the wild type SeADO. The Perez et al., 2011; Moss et al., 2016).
author suggested that this improvement was due to the increase of Almost all filamentous cyanobacteria that contain the OLS pathway
protein solubility and removal of the side chain that blocks substrate have the FAAL-ACP1 and KS-AT-KR-ACP2-ST-TE systems encoded in
entry (Wang et al., 2017). two open reading frames except Leptolyngbya sp. PCC 7376 (unicellular
Investigation of ADO substrate profiling revealed that ADOs can use and baeocystis cyanobacteria), in which all enzyme genes are encoded
short-, medium-, and long-chain fatty aldehydes (C4–C18) as substrates in only one reading frame.
(Andre et al., 2013; Bao et al., 2016; Khara et al., 2013). This indicates

8
J. Jaroensuk, et al. Journal of Biotechnology 309 (2020) 1–19

Fig. 6. The crystal structure of SeADO (PDB ID code 4RC5) (Jia et al., 2015). Residues close to the di-iron center (Y39, R62, Y122, D143) are shown, as well as a key
residue (W178) located on the protein surface and close to the active site, and residues located close to the hydrophobic tail of the aldehyde substrate (Y21 and C70).

3.3.2. Olefin-forming fatty acid decarboxylase (OleT) utilizing H2O2 (peroxide shunt pathway) to generate a ferric
Olefin-forming fatty acid decarboxylase (OleT) is an attractive en- (FeIII)‐hydroperoxo (compound 0) which is then protonated and dehy-
zyme for various applications and has drawn great interest due to its drated to yield a high energy ferryl (FeIV)-oxo species (compound I).
ability to transform free fatty acids to 1-alkenes in a single-step reac- Compound I was observed in single turnover reactions of OleTJE using
tion. OleT found in Jeotgalicoccus sp, Bacillus subtilis, Corynebacterium stopped‐flow absorption spectroscopy (Grant et al., 2015) ; (V) Com-
efficiens, Kocuria rhizophila, and Methylobacterium populi belong to the pound I abstracts a hydrogen atom from the β-carbon position of the
cytochrome P450 CYP152 family and are known to catalyze the dec- fatty acid, generating a substrate radical and the ferryl (FeIV)-hydroxo
arboxylation of medium- to long-chain (C12-C20) fatty acids into 1- species (compound II). (VI) Compound II in OleTJE undergoes two
alkenes (C11-C19) using H2O2 as an oxidant and electron donor possible pathways which are decarboxylation of the fatty acid to form
(Belcher et al., 2014; Matthews et al., 2017; Munro et al., 2018; Rude 1-alkene (g–h) or hydroxylation the substrate at the α‐ or β‐position
et al., 2011; Zachos et al., 2015). Moreover, OleT also catalyzes α- and (i–j).
ß- hydroxylation of fatty acids as a side reaction (Lee et al., 2003;
Matsunaga et al., 2000; Rude et al., 2011; Shoji et al., 2007).
3.3.3. Nonheme mononuclear iron oxidase UndA and UndB
Besides direct addition of H2O2 to the reaction to support OleTJE
UndA and UndB are two major enzymes responsible for the bio-
production of 1-alkenes from fatty acids, the use of light or an H2O2-
synthesis of 1-alkenes in the Pseudomonas species (Rui et al., 2015,
forming enzyme to generate H2O2 for OleTJE has also been explored
2014). Biochemical analysis of UndA demonstrated that UndA requires
(Matthews et al., 2017; Zachos et al., 2015). Fusion of an H2O2-gen-
FeII as a cofactor and O2 as an activator for catalyzing the decarbox-
erating enzyme, alditol oxidase (AldO) from Streptomyces coelicolor,
ylation of medium-chain fatty acids (C10-C14) to generate 1-alkene
with OleTJE has been employed to continuously supply H2O2 for the
with release of CO2 (Rui et al., 2014). Biochemical characterization of
oxidative conversion of fatty acids into 1-alkenes (Matthews et al.,
UndA showed that the stoichiometry of O2 binding to UndA is
2017). This approach was used successfully for the production of 1-
equivalent to the formation of 1-alkene. The results of an in vitro assay
alkenes with a 98 % conversion yield of myristic acid (C14:0) which is
using lauric acid (C12:0) as a substrate also showed that UndA can only
higher than that obtained by direct addition of H2O2 at the start of the
catalyze one turnover of lauric acid (C12:0) to form 1-undecene. Ad-
reaction (87 % conversion yield).
dition of reductants including ascorbate, PMS/NADH, and PQQ/cy-
Recently, Fang et al. (2017) also showed that in addition to H2O2,
steine, could improve the turnover number of UndA in vitro to about 4
OleTJE can also use other redox partner proteins/NAD(P)H-utilizing
turnovers per hour (Rui et al., 2014).
systems to generate 1-alkenes from fatty acids with a 94.4 % yield. The
The results of substrate specificity analyses using lauric acid (LA)
most efficient redox partner protein for 1-alkene production reported in
derivatives which are 2-hydroxydodecanoic acid (AHDA), [α,α-D2]LA,
this study was Fd from Corynebacterium glutamicum and FNR from S.
[D23]LA, 3-hydroxydodecanoic acid (BHDA), and 2,3-dodecanoic acid
elongatus.
(DEA), showed that (I) UndA could convert α-hydroxy fatty acids into
The mechanism of OleT catalysis of the decarboxylation of fatty
1-alkenes but could not use ß-hydroxy fatty acids as a substrate, (II) 1-
acids into 1-alkenes or α- or ß-OH fatty acids has been proposed (Fig. 7)
undecene produced from [α,α-D2]LA retained both deuterium atoms,
(Belcher et al., 2014; Lee et al., 2003; Munro et al., 2018; Rude et al.,
and (III) the UndA assay with [D23]LA resulted in the formation of
2011). The reaction of OleT consists of six steps: (I) Fatty acids bind to
[D22]1-undecene (Rui et al., 2014). Based on this biochemical analysis,
the iron center by displacing the axial water molecule and converting
Rui et al. (2014) proposed a mechanism to describe UndA catalysis of
the heme iron from a low spin to a high spin state (Sligar and Gunsalus,
the decarboxylation of fatty acids into 1-alkenes. The reaction of UndA
1979); (II-IV) OleT bypasses the O2-redox partner system (a–d) by
consists of three steps: (I) fatty acid binds to the mononuclear iron (FeII)

9
J. Jaroensuk, et al. Journal of Biotechnology 309 (2020) 1–19

Fig. 7. Proposed mechanism of OleT catalysis of the decarboxylation of fatty acids into 1-alkenes or α- or ß−OH fatty acids (Belcher et al., 2014; Lee et al., 2003;
Munro et al., 2018; Rude et al., 2011).

Fig. 8. Proposed mechanism of UndA synthesis of an alkene with a terminal double bond (Rui et al., 2014).

center at the active site of UndA and forms intermediate 1 (FeII-fatty ferredoxin reductase (EcFd-EcFNR), flavodoxin-ferredoxin reductase
acid), (II-III) the Fe(III)-superoxide complex formed at the active site of (EcFld-EcFNR)), and from Pseudomonas putida (e.g. putidaredoxin-puti-
UndA following O2 binding abstracts a hydrogen from the ß-carbon of daredoxin reductase (Pd-PNR)) with UndB also increased the produc-
the fatty acid, followed by one electron transfer and scission of the tion of 1-alkenes in engineered E. coli by approximately 1.5-, 1.75-, and
carbon-carbon bond, resulting in formation of 1-undecene, CO2 and 2.25-fold as compared to the engineered strain harboring only UndB,
H2O (Fig. 8) (Rui et al., 2014). Structural analysis of UndA in complex respectively (Zhou et al., 2018a).
with 3-hydroxydodecanoic acid (BHDA) also supports the mechanism
proposed above because the position of the ß-hydrogen of the fatty acid 3.3.4. Fatty acid photodecarboxylase (FAP)
is ideal for abstraction, as the distance between the ß-hydroxyl oxygen FAP is a flavoenzyme found in Chlorella green algae that possesses
of BHDA and the distal oxygen atom is only 2.45°A (Rui et al., 2014). the ability to produce hydrocarbons through the photodecarboxylation
UndB, a membrane-bound desaturase-like enzyme, catalyzes the of fatty acids (Sorigue et al., 2017; Yu et al., 2019). This enzyme needs
formation of 1-alkene from short- to medium-chain fatty acid (C6-C16) photons from blue light (400−520 nm) to activate its FAD cofactor for
in the same manner as UndA. UndB is only distributed in a few species catalysis as shown in Fig. 9 (Huijbers et al., 2018; Sorigue et al., 2017;
of Acinetobacter and Pseudomonas, such as P. fluorescens and P. mendo- Sorigué et al., 2016; Zhang et al., 2019). Several studies have shown
cina. Overexpression in E. coli of UndB with UcFatB2, a thioesterase that FAP exhibits turnover numbers (TON) in the range of thousands
which specifically hydrolyzes the lauryl chain from ACP resulted in a (Huijbers et al., 2018; Zhang et al., 2019), suggesting that use of FAP
100-fold improvement in 1-undecene production as compared to a for large-scale production of alka(e)nes from fatty acids may be feasible
strain in which UndB was overexpressed alone. The results indicate that (Hoeven et al., 2019; Scrutton, 2017).
free fatty acids are the preferred substrates of UndB. Additionally, co- The mechanism of FAP catalysis of the decarboxylation of fatty
expression of electron transfer systems from E. coli (e.g. ferredoxin- acids to Cn-1 alka(e)nes has been proposed (Fig. 9) (Huijbers et al.,

10
J. Jaroensuk, et al. Journal of Biotechnology 309 (2020) 1–19

Fig. 9. The reaction mechanism of FAP conversion of a fatty acid into alka(e)ne starts by a base abstraction of a fatty acid to result in a carboxylate anion and FAD
(RCOO−-FAD). The di-radical complex (RCOO●…●-FAD) is produced after an excited singlet FAD* abstracts one electron from the carboxylate substrate. Finally, the
reduction of the alkyl radical results in release of the alka(e)ne and CO2 (Huijbers et al., 2018; Sorigue et al., 2017; Zhang et al., 2019).

2018; Sorigue et al., 2017; Zhang et al., 2019). The mechanism consists 4. Metabolic Engineering or cascade reactions for the synthesis of
of three steps: (I) a proton from the fatty acid is abstracted by an un- hydrocarbons
known base (B−) to generate the carboxylate anion that will then form
an intermediate complex with FAD (RCOO−-FAD). Theoretical studies The overall diagram describing the engineered metabolic pathways
from density functional theory (DFT) and time dependent-DFT (TD- for conversion of fatty acids to hydrocarbons is in Fig. 10.
DFT) revealed that coordination between the carboxylate group and the
N5 site of FAD is required to initiate the activity of FAP (Zhang et al.,
4.1. Carboxylic acid reductase-Aldehyde deformylating oxygenase (CAR-
2019). In steps II-III, an excited singlet state, FAD*, generated from the
ADO)
photo excitation, abstracts one electron from the carboxylate substrate
to yield a di-radical complex (RCOO●…●-FAD) which then rapidly
The combined use of CAR and ADO has been employed in various
converts to an alky radical, and later an alka(e)ne. The reduction of an
engineered pathways for the in vivo production of aliphatic hydro-
alkyl radical to an alka(e)ne (step III) has been speculated to occur via
carbons. This system was first reported in 2013 by Akhtar et al. (2013)
two mechanisms (Sorigue et al., 2017): (a) the alkyl radical receives one
in which the CAR enzyme from Mycobacterium marinum (MmCAR) and
electron from FAD●- and one proton from a proton donor via a proton-
the cyanobacterial ADO from Prochlorococcus marinus (PmADO) were
coupled electron transfer mechanism to yield B− and release of alka(e)
co-expressed in E. coli along with two additional enzymes, namely E.
ne; (b) the alkyl radical receives one hydrogen from BH, followed by
coli thioesterase A (TesA) which catalyzes conversion of fatty acyl-ACP/
one electron transfer from FAD●- to B●, generating B− and releasing
CoA to free fatty acids and Bacillus subtilis Sfp as a PPT to activate CAR
the alka(e)ne (Sorigue et al., 2017).
(Beld et al., 2014). In the presence of exogenous fatty acid, the culture
The substrate specificity of FAP has been investigated (Huijbers
of CAR-ADO containing cells produced a mixture of medium-chain
et al., 2018; Zhang et al., 2019). FAP from Chlorella variabilis NC64A
hydrocarbons at an approximate concentration of 2 mg/L. It was,
(CvFAP) exhibits the ability to utilize C1-C20 fatty acids as substrates.
however, suggested that such poor yield was likely the result of very
However, CvFAP showed greater activities towards long-chain fatty
low turnover rates of ADO and accumulation of cellular toxicants
acids (C > 14) than short- and medium-chain fatty acids. It has been
(Andre et al., 2013). This was likely due to the fact that aldehyde re-
reported that a 90–96 % conversion of C16:0, C17:0, C18:0, and C20
ductases and dehydrogenases can also compete with ADO for fatty al-
substrates could be obtained after performing the photoenzymatic re-
dehyde intermediates (Akhtar et al., 2013; Kallio et al., 2014).
action (blue light intensity = 13.7 μEL−1s−1) for 14 h in 100 mM Tris-
The MmCAR-PmADO pathway has also been utilized for the pro-
HCl (pH 8.5) containing 30 mM fatty acid substrate, 30 % DMSO, and
duction of short-chain alkanes such as propane in E. coli (Kallio et al.,
6.0 μM of crude extract CvFAP (Huijbers et al., 2018). Moreover,
2014; Menon et al., 2015). The principal concept for biosynthesis of
Huijbers et al. (2018) also observed that the crude extract CvFAP is
propane by the MmCAR-PmADO pathway is similar to that of the
much more active and stable than the purified CvFAP.
medium-chain alkanes previously mentioned (Beld et al., 2014). Im-
provements of the system were done through (I) the substitution of E.

11
J. Jaroensuk, et al. Journal of Biotechnology 309 (2020) 1–19

Fig. 10. Overview of engineered metabolic pathways for hydrocarbon biosynthesis from fatty acids. The pink panel represents fatty acid biosynthesis, the central
metabolism for the biosynthesis of hydrocarbon using endogenous fatty acid. The gray panel shows the pathway related to alkane synthesis. The yellow panel shows
the pathway involved in alkene synthesis. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article).

coli TesA with Bacteroides fragilis Tes4 which was reported to have reaction is widely believed to be the rate-limiting step for bio-alkane
higher specificity towards C4 butyryl-ACP (Kallio et al., 2014); (II) the production (Patrikainen et al., 2017; Warui et al., 2015).
substitution of wild-type PmADO with A134 F PmADO that has a high Another feasible cell factory explored for hydrocarbon formation
affinity and activity towards butyraldehyde (Khara et al., 2013; was Saccharomyces cerevisiae. According to Zhou et al. (2016b), the
Sheppard et al., 2016); (III) the addition of redox partner proteins yeast strain was first engineered to be capable of accumulating high
(Synechocystis sp. PCC6803 Fd (SynPCC608Fd) and E. coli FNR (EcFNR)) concentrations of free fatty acids (FFAs) prior to equipping the cell with
and E. coli catalase (KatE); (IV) deletion of ahr and yqhD which encode the alkane-producing system including E. coli Fd (EcFd), EcFNR,
the enzymes that can competitively consume butyraldehyde, the im- MmCAR, and NpADO. This engineered yeast strain was found to give
mediate precursor of propane synthesis. This strategy led to a remark- alkane of up to 0.82 mg/L, which is still very low. The yield improve-
able improvement in propane production (32 mg/L) after 19-h culti- ment in yeast could be achieved by peroxisomal compartmentalization
vation (Kallio et al., 2014). of the biosynthetic routes in order to minimize the cytosolic toxicity of
Due to the fact that the microbial platforms for propane biosynthesis H2O2 and spatially sequester the synthetic enzymes of interest away
rely on the flux of butyryl-ACP (precursor) from fatty acid biosynthesis, from competing reactions (Zhou et al., 2016a; Zhu et al., 2017). This
alternative pathways to redirect the precursors towards formation of strategy resulted in production of alkanes at a level of about 4 mg/L by
propane have been explored. Menon et al. (2015) designed a set of the engineered yeast strain with the expression of S. elongatus Fd (SeFd),
alternative CoA-dependent butyraldehyde synthetic routes which uti- S. elongatus FNR (SeFNR), MmCAR, and SeADO in peroxisomes (Zhou
lize enzymes acetyl-CoA acetyltransferase (AtoB) and acetoacetyl-CoA et al., 2016a).
synthase (NphT7) encoded by the genes atoB and nphT7, respectively, Altogether, it is clear that further improvement is needed for en-
to be overexpressed in E. coli BL21 (DE3). Coupling of MmCAR-A134 F hancing the production efficiency of hydrocarbon biosynthesis while
PmADO-SynPCC608Fd-EcFNR with AtoB and NphT7 was able to pro- circumventing the potential bottlenecks, in either yeast or bacteria, for
duce propane with a titer of 3.40 ± 0.19 mg/L over 12 h of cultivation production of alkanes by the CAR-ADO systems. More investigation of
(Menon et al., 2015). the CAR-ADO platform in both prokaryotic and eukaryotic hosts as well
More recently, the CAR-ADO pathway has been investigated in as a better understanding of ADO catalysis can improve the perfor-
order to identify the best combination of enzymes for the production of mance of these engineered cells for hydrocarbon production.
alkanes from exogenous fatty acids in E. coli. ADOs from various strains
of cyanobacteria were combined with MmCAR, SynPCC608Fd and
4.2. Acyl-[acyl carrier protein (ACP)] reductase-Aldehyde deformylating
EcFNR for C3-C7 alkane biosynthesis (Patrikainen et al., 2017). The
oxygenase (AAR-ADO)
highest yield was obtained with the engineered system bearing NpADO,
giving a heptane yield of around 35 % from exogenous fatty acid. It can
The AAR-ADO system has been investigated in several studies for
thus be inferred that selection of an efficient ADO is one of the crucial
the production of long-chain alkanes. With heterologous expression of
factors to achieving successful production yields of alkane because the
SeAAR and NpADO in E. coli, the alkane titer (tridecane, pentadecene,

12
J. Jaroensuk, et al. Journal of Biotechnology 309 (2020) 1–19

pentadecane and heptadecane) was able to reach 300 mg/L in which NAD(P)H, an important reductant of alkane biosynthesis (Fig. 10)
more than 80 % of the product is secreted outside the cell (Schirmer (Jaroensuk et al., 2019). Based on this strategy, we managed to produce
et al., 2010). The possibility of introducing multiple copies of aar-ado alkane with a 50 % yield from exogenous fatty acid which is the best
into the genome of cyanobacterial host strain was also explored. Wang yield to date. Therefore, this approach should be useful for future
et al. (2013) succeeded in obtaining the engineered Synechocystis sp. production of fatty acid-derived chemicals.
PCC6803 strain with extra copies of aar-ado in both slr0168 (unknown
function) and slr1556 (2-hydroxyacid dehydrogenase gene, ddh) loci. 4.4. Terminal alkene synthesis
The engineered Synechocystis sp. PCC6803 resulted in production of
alka(e)ne with titer of 26 mg/L from a 50 ml culture in a photo- OleTJE, UndA, and UndB have been utilized for producing terminal
bioreactor. This level of alkane production is approximately eight-fold alkenes in many engineered cell factories such as E. coli and S. cerevisiae
higher than the production levels attained with the wild-type strain, (Zhou et al., 2016a, a). In E. coli, Liu et al. (2014b) engineered E. coli by
suggesting that this approach may be used for future hydrocarbon deletion of the acyl-CoA synthetase gene (fadD) and overexpression of
production by genetically engineered hosts. acyl-CoA carboxylase (ACC), thioesterase A (TesA) and OleTJE or OleTJE
Other plausible genetic manipulations, particularly those which fused with a CYP450 reductase domain from Rhodococcus sp (RhFRED),
supply other additional associated genes responsible for driving the and used the engineered cells for production of 1-alkene. Interestingly,
metabolic flux towards alka(e)ne biosynthesis, have often been per- the engineered E. coli harboring OleTJE produced total 1-alkene
formed following the establishment of AAR-ADO in these cells. For (96.7 mg/mL) more than that of the engineered E. coli harboring OleTJE
example, Coursolle et al. (2015) revealed that the co-expression of type- fused with RhFRED (6.0 mg/mL). The author suggested that high pro-
I fatty acid synthase (FAS) from Corynebacterium ammoniagenes with duction of alkene in E. coli harboring OleTJE was due to the ability of
SeAAR-SeADO in the BL21(DE3) enhanced the total titer of mixed long- OleTJE to use electron transfer protein in E. coli (i.e. Fd & FNR) to drive
chain alkane/alcohol to roughly 100 mg/L, in which the pentadecane catalysis efficiently.
content was 57 mg/L. This yield is higher than when only SeAAR- In yeast, Zhou et al. (2018a) demonstrated the overexpression of
SeADO (∼8.15 mg/L) was expressed (Coursolle et al., 2015) (Table 1). OleTJE, P. putida UndA (PpUndA), P. mendocina UndB (PmUndB), and P.
One of the most notable improvements for long-chain alkane bio- fluorescens (PfUndB) in a fatty acid-overproducing yeast strain (S. cer-
synthesis in E. coli via the AAR-ADO pathway was recently reported by evisiae YJZ06) for production of 1-alkene. All the engineered yeast
Fatma et al. (2018) In this study, the flux towards fatty acid synthesis strains were able to produce 1-alkene, however with a very low titer
routes in E. coli was enhanced by upregulating the gene encoding for (∼1−6 mg/L). Among the four engineered yeast strains, the strain
glucose-6-phosphate dehydrogenase (zwf) to boost the accumulation of harboring PfUndB produced the highest titer of 1-alkenes. To improve
intracellular NADPH. Meanwhile, six sets of genes including the Entner- the production, various electron transfer systems were co-expressed
Doudoroff pathway (phosphogluconate dehydratase, edd), gluconeo- with PfUndB in S. cerevisiae to enhance the reduction of FeIII to FeII, an
genesis pathway (phosphoenolpyruvate synthase, ppsA), fermentative important cofactor of UndB. The Pd-PNR showed the highest im-
pathway (D-lactate dehydrogenase, ldhA; pyruvate oxidase, poxB), provement of PfUndB activity. The 1-alkene production was improved
glyoxylate pathway (isocitrate lyase, aceA) and phospholipid pathway by 50 % compared to the strain expressing only PfUndB. In addition, a
(acyl-ACP:phosphate transacylase) were deleted to remove the path- long-chain fatty acid transporter (FATP1) from Homo sapiens was en-
ways competing for glucose. The engineered cells resulted in an alkane gineered into S. cerevisiae. All efforts led to the production of 1-alkene
titer of 425 mg/L. The production was elevated to 2.54 g/L after 72-h with a titer of 35 mg/L, of which, 80 % was secreted into the medium
fed-batch cultivation in a 5 L fermenter using glucose as substrate (Zhou et al., 2018a).
(500 g/L). The alkane titer from this work has surpassed the previously Terminal alkene synthesis by an in vitro enzymatic cascade has also
highest titer (1.31 g/L) obtained from a 2.5 L fed-batch culture of an been explored as an alternative strategy. Dennig et al. (2015) employed
engineered strain with AAR-ADO overexpression and an electron OleTJE and Pd-PNR, an effective redox partner of OleTJE for production
transfer system from S. elongatus (Cao et al., 2016). of 1-alkene. By this system, they acquired 0.93 g/L of C17 alkene from
Combination of AAR-ADO expression in yeast resulted in much octadecanoic acid as a starting substrate and turnover numbers of more
lower efficiency of hydrocarbon generation than the CAR-ADO pathway than 2000 per 8 h. Yan et al. (2015) demonstrated a one-pot reaction
(Zhou et al., 2016a, b). However, alkane synthesis by strains bearing for the synthesis of 1-alkene from triacylglycerols (TAGs) using lipase
either AAR or CAR with various ADO and electron-donating modules from Thermomyces lanuginosus (Tllipase) and OleTJE. Although 1-al-
was improved with the compartmentalization approach (Zhou et al., kenes were produced from the TAGs, the yield and reaction rate were
2016a, b). Nevertheless, it could be seen that the use of yeast cells for very low. Recently, Li et al. (2019) developed a new one-pot process for
overexpression of AAR-ADO for hydrocarbon production is not as effi- production of 1-alkenes from TAGs by immobilizing the dockerin-taq
cient as the use of bacterial hosts. enzymes (Tllipase-D1 and P450 OleTJE-D2) on the cohesion-cellulose
binding module. The results showed that the co-immobilization of
4.3. Fatty acyl-CoA reductase-Aldehyde deformylating oxygenase (ACR- OleTJE with Tllipase on the cellulose matrix enhanced the reaction rate
ADO) and conversion of TAGs to 1-alkenes. The reaction rate and yield in-
creased by 9-fold and 2.8-fold respectively, when compared to the re-
Unlike the previous two systems mentioned, overexpression of ACR action using a mixture of free enzymes. The most efficient cascade for 1-
and ADO to generate engineered metabolic cells for production of al- alkene production was comprised of a 1:2 ratio of Tllipase to OleTJE. In
kanes has only been recently investigated (Akhtar et al., 2013; addition, the stability and recyclability of the immobilized enzyme
Jaroensuk et al., 2019). In the past, overexpression of ACR has been complexes were better than those of the free enzyme mixtures. The
used for production of other fatty acid-derived chemicals such as fatty activities of the immobilized enzyme complexes remained around 89 %
alcohol and wax esters (Lehtinen et al., 2018; Santala et al., 2011, after reuse for 10 times.
2014). Recently, our team has successfully developed an artificial me-
tabolic pathway comprised of ACR, ADO, Fd, FNR, and XaFDH for the 4.5. Fatty acid photodecarboxylase (FAP)
synthesis of hydrocarbons from exogenous fatty acid. A key invention of
our system is the improvement ACR-ADO mediated alkane production Overexpression of FAP to generate the metabolically engineered
by addition of formate dehydrogenase (FDH) activity. FDH from Xan- cells for production of alkane has only been recently investigated
thobacter sp. 91 (XaFDH was added into the metabolic pathway so that (Bruder et al., 2019; Hoeven et al., 2019). Hoeven et al. (2019) de-
it can convert formate, an unused co-product of ADO reaction, into monstrated the utilization of a variant CvFAP G462 V for production of

13
J. Jaroensuk, et al.

Table 1
Scale-up production of hydrocarbon compounds by engineered microbial strains.
Ref. Fermentation mode (bioreactor Time (h) Culture media Carbon Feeding solution Microorganism Product titer (g/L) Productivity (g/L/h)
size in liter) source

(Kim et al., Fed-batch mode (5 or 6.6 L); 101.3 Minimal medium 40 g/L 800 g/L glucose, 5 g/L Rhodococcus opacus ROPA2 5.20 (C12, C14-C19, C21, C24, 0.051
2019) initial working volume (2 L); glucose MgSO4•7H2O (100 mL C27 alkanes)
final volume (2.48 L) feeding)
(Fatma et al., Fed-batch mode (5 L); initial 84 Modified M9 medium 20 g/L 200 g/L yeast extract, Escherichia coli strain ZFM8 2.50 (C13, C15, C17 alkanes; 0.031
2018) working volume (3 L) glucose 500 g/L glucose, 12.47 g/L C15, C17 alkenes)
MgSO4
(100 mL feeding)
(Wang et al., Fed-batch mode (5 L); initial 48 Modified M9 medium 30 g/L 2−3 g/L/h glucose E. coli BL21 (DE3) with 1.009 (C15, C17 alkanes; C17 0.021
2018) working volume (n.a.) glucose overexpression of fadR-ΔyqhD alkene)

14
(Choi and Lee, Fed-batch mode (6.6 L); initial n.a. MR medium 20 g/L 400 g/L glucose, 14 g/L E. coli w3110 strain GAS3 with 0.580 (C9, C12, C13, C14 n.a.
2013) working volume (2 L) glucose MgSO4•7H2O (100 mL tesA(L109 P) alkanes)
feeding) E. coli w3110 with tesA 0.804 (C13, C15 alkanes) n.a.
(Xin et al., 2017) Batch mode (10 L); initial 120 Heavy oil producing 160 g/L – Aureobasidium melanogenum 9-1 43.0 (89.8 % alkane; 10.19 % 0.358 (product);
working volume (7 L) medium inulin transformant 88 (inulinase C16, C18 fatty acids) 0.354 (alkane)
overexpression)
A. melanogenum 9-1 38.24 (87.12 % C20-C34 0.319 (product);
alkanes; 12.88 % C16, C18 0.278 (alkane)
fatty acids)
(Cao et al., Batch mode (5 L); initial working 50.5 (40.5 h after Mineral medium 15 g/L 500 g/L glycerol, 200 g/L E. coli BL21 (DE3) with plasmid 1.31 (C15 alkane; C15, C17 0.032
2016) volume (2.5 L) induction) glycerol yeast extract, and 2.47 g/L YX10+SeFd/Fnr alkenes)
MgSO4
(Chen et al., Fed-batch mode (5 L); initial 144 YPD medium + G418 30 g/L 200 g/L glucose (150 mL S. cerevisiae strain BY22 0.0037 (C11, C13, C15, C17, 0.000026
2015) working volume (1 L) medium glucose feeding) C19 alkenes)
(Liu et al., Batch mode (10 L); initial 168 Heavy oil producing 120 g/L – A. melanogenum (native strain) 32.5 g/L heavy oil (66.15 % 0.128
2014a) working volume (6 L) medium glucose C13, C15, C24, C26, C27, C28,
C44 alkanes; fatty acids 26.38
%)
Journal of Biotechnology 309 (2020) 1–19
J. Jaroensuk, et al. Journal of Biotechnology 309 (2020) 1–19

propane in E. coli, Halomonas sp, and Synechocystis sp. The engineered In addition to E. coli, other microbial hosts were also explored to
E. coli harboring CvFAP G462 V was found to produce propane produce hydrocarbon on a larger fermentation scale. For example, an
(48.31 + 2.66 mg/L) from 10 mM butyric acid within 18 h. The titer of engineered S. cerevisiae strain BY22 was reported to produce about
propane was further improved by adding ethyl acetoacetate into the 3.7 mg/L of alkene, with a relatively low productivity rate of
medium to activate a butyrate uptake-transporter, atoE. The titer of 0.000026 g/L/h, even after 144 h of fed-batch cultivation using glucose
propane obtained by this approach was 97.1 + 10.3 mg/L, the highest as carbon source (Chen et al., 2015). Investigations using exogenous
titer reported to date for biopropane production using engineered E. coli fatty acid as a sole carbon source for alkane biosynthesis by engineered
(Hoeven et al., 2019; Kallio et al., 2014; Menon et al., 2015). S. cerevisiae were later explored (Foo et al., 2017). The fermentation of
The significant improvement of propane production was obtained an engineered S. cerevisiae strain (Δfaa1Δfaa4) with overexpressed
when CvFAP G462 V was overexpressed in Halomonas sp. (Hoeven SeADO and the enzyme which can produce Cn-1 aldehydes from fatty
et al., 2019). Under non-sterile conditions, the engineered Halomonas acids (α-dioxygenase (α-DOX)) from Oryza sativa in a medium con-
can produce propane at a level of 157 +/- 17 mg/L from 80 mM butyric taining 200 mg/L exogenous fatty acid (C12:0-C18:0) as the sole carbon
acid within 48 h, the highest titer reported to date for biopropane source was reported to produce 28.7, 304.0, and 5.1 mg/L of dodecane,
production using an engineered microbe (Hoeven et al., 2019; Kallio tetradecane, and hexadecane at pH 7.0, after 48 h, respectively (Foo
et al., 2014; Menon et al., 2015). To expand the use of CvFAP for et al., 2017). Nevertheless, the alkane titers were still low. This has led
production of propane from CO2, CvFAP was expressed in Synechocystis to exploration on other microorganisms for biosynthesis and conversion
sp. The titer of propane produced from this engineered cyanobacterium of carbon sources to hydrocarbons. The bacterium R. opacus has been
was reported at 12.2 +/- 2.6 mg propane/g cells/day (Hoeven et al., used as a host candidate for hydrocarbon synthesis. The engineered R.
2019). opacus strain (ΔfadEΔalk-1) with overexpressed of SeADO and enzymes
Hydrocarbons synthesis by in vitro enzymatic cascade has also been for the production of free fatty acids (TAG lipase (LPD05381) from R.
explored. A full length CvFAP could not be overexpressed. Only the opacus and lipase chaperon (LSF) from Burkholderia cepacian) and en-
CvFAP variant lacking a chloroplast-targeting sequence could be over- zymes for the production of fatty aldehydes (acyl-CoA reductase (ACR)
expressed in soluble form (Huijbers et al., 2018; Zhang et al., 2019). A from Clostridium kluyveri) was able to produce 5.20 g/L of alkane with
two-step reaction for the synthesis of alkanes from triglycerides using the productivity rate of 0.051 g/L/h when glucose was used as a sole
lipase from Candida rugosa (Crlipase) and the CvFAP variant could be carbon source under a fed-batch process (Kim et al., 2019).
used to synthesize heptadec-8-ene from triolein with more than 80 % While significant efforts have been devoted to engineer microbial
yield and turnover number (TON = molproduct × molFAP −1) of 8280 strains for greater superiority in forming hydrocarbon products, the
(reaction time used of 20 h) (Huijbers et al., 2018). In another recent culture of some native strains, including Aureobasidium melanogenum
study on hydrocarbons synthesis, a decoy molecule was used to fill up strain P5, an oleaginous fungal endophyte found to be associated with a
part of the substrate tunnel of CvFAP, resulting in expansion of sub- mangrove species, has already been observed to give process yields and
strate scope to include carboxylic acids with straight-chain, branched, productivities acceptable for larger scale production. The strain
and unsaturated structures and to increase the rate of decarboxylation P5 grown in a 6 L initial working volume of batch culture yielded a
(Zhang et al., 2019). In addition to hydrocarbon synthesis, CvFAP was mixture of n-alkanes and fatty acids (32.5 g/L), with C13, C15, C24,
also used to synthesize chiral α-hydroxy/amino carboxylic acids (Xu C26, C28, and C44 alkanes adding up to a total yield of 66 % (21.5 g/L)
et al., 2019). The results demonstrated that a single residue substitution with a productivity rate of 0.128 g/L/h, which was higher than those of
(G426Y) in the substrate channel of CvFAP can increase stereo- the engineered hosts in several other studies (Liu et al., 2014a). How-
selectivity of this reaction significantly (ee up to 99 %). ever, the highest production performance to date has been found in a
different strain of A. melanogenum 9-1 with overexpressed inulinase,
5. Large scale production giving a yield of 43.0 g/L of heavy oil containing 89.8 % alkane at a
productivity rate of 0.354 g/L/h in a batch process with 7 L starting
Despite there being many types of native microorganisms that are volume fed with inulin as the sole carbon source (Xin et al., 2017).
capable of synthesizing hydrocarbons, their production yields are all For a scale-up cultivation for hydrocarbon production, it can be seen
low and impossible to employ in industrial processes. Therefore, most that a wide range of carbon sources can be chosen as well as different
studies of scaling up fermentation processes have only been focused on fermentation modes. Possible feeding substrates include simple sugars
the metabolically engineered or genetically modified (GM) microbial (e.g. glucose, fructose, galactose and lactose), fatty acids or even bio-
hosts, e.g. E. coli, Rhodococcus opacus, S. cerevisiae and Yarrowia lipoly- mass such as lignocellulosic materials (Grim et al., 2019; Hoeven et al.,
tica (Jimenez-Diaz et al., 2017; Schirmer et al., 2010, 2014) Redirecting 2019; Kohli et al., 2019; Luo et al., 2019; Ng et al., 2019; Toogood and
the host system to produce a higher titer of products via metabolic Scrutton, 2018; Zhou et al., 2018b). Although direct supplementation of
engineering is generally perceived as one of the major approaches to exogenous fatty acids has not been reported on such scale, some studies
achieve successful production of alka(e)nes for industrial use. In 2013, attempted to construct the engineered strains with upregulation of
the engineered E. coli W3310 was reported to produce alkane up to enzymes such as 3-oxoacyl-ACP synthase (FabH), which is responsible
0.804 g/L from glucose using fed-batch fermentation (Choi and Lee, for the production of endogenous fatty acid to be subsequently con-
2013). Using the same mode of cultivation, the titers of alkane pro- verted to hydrocarbon in vivo (Choi and Lee, 2013). As an alternative
duced by the engineered E. coli strain BL21 (DE3) with overexpression feedstock, glycerol, a by-product derived from saponification or trans-
of aar-ado-fadR and deletion of yqhD and the E. coli strain DH5α with esterification of triglycerides, was also explored as another source of
overexpression of aar-ado-zwf and deletion of edd, ppsA, ldhA, aceA, carbon. Culture of the engineered E. coli BL21 (DE3) with alkane bio-
poxB and plsX reached 1.009 and 2.50 g/L, with productivity rates of synthetic genes using a 2.5 L fed-batch process in glycerol feed for
0.021 and 0.031 g/L/h respectively (Fatma et al., 2018; Wang et al., 50.5 h yielded a mixture of C15 alkane and C15 and C17 alkenes with a
2018). The low production titers are thought to be due to the fact that yield of 1.31 g/L and a productivity rate of 0.032 g/L/h (Cao et al.,
the glucose substrate fed into the culture is predominantly catabolized 2016). In terms of fermentation mode, either batch or fed-batch process
and used in other pathways to generate energy, biomass and other is commonly implemented to grow producer strains when scaling up.
metabolites besides hydrocarbon (Kim et al., 2019). Further investiga- However, continuous operation for manufacturing hydrocarbon com-
tion and evaluation of large-scale microbial fermentation for an im- pounds has not been reported. One underlying reason could be due in
proved and sustainable hydrocarbon-producing system are required to part to the low yield of hydrocarbon production. More optimization of
make the systems into economically viable modes of production of yield improvement is needed so this technology can reach the level
hydrocarbons. which is needed for economy of scale.

15
J. Jaroensuk, et al. Journal of Biotechnology 309 (2020) 1–19

6. Future perspectives References

Based on the current technologies available (Sections 2–5), it can be Akhtar, M.K., Turner, N.J., Jones, P.R., 2013. Carboxylic acid reductase is a versatile
seen that various types of enzymatic cascades and metabolic engineered enzyme for the conversion of fatty acids into fuels and chemical commodities. Proc.
Natl. Acad. Sci. U. S. A. 110, 87–92. https://doi.org/10.1073/pnas.1216516110.
cells are capable of enabling de novo synthesis of alkanes and alkenes Amelia, T., Govindasamy, S., Tamothran, A., Vigneswari, S., Bhubalan, K., 2019.
from bio-based feedstocks such as sugars and fatty acids (Fig. 10). Applications of PHA in Agriculture. Biotechnological Applications of
As the current and future trends of industries will require all pro- Polyhydroxyalkanoates. Springer, Singapore, pp. 347–361. https://doi.org/10.1007/
978-981-13-3759-8_13.
duction to meet UN 17 sustainable development goals and any in- Andre, C., Kim, S.W., Yu, X.H., Shanklin, J., 2013. Fusing catalase to an alkane-producing
dustries with high emission of greenhouse gas will be required to pay enzyme maintains enzymatic activity by converting the inhibitory byproduct H2O2 to
carbon tax (Ramseur and Leggett, 2019), enzymatic and cell-based the cosubstrate O2. Proc. Natl. Acad. Sci. U. S. A. 110, 3191–3196. https://doi.org/
10.1073/pnas.1218769110.
methods serve as beautiful alternative technologies for hydrocarbon Bao, L., Li, J.J., Jia, C., Li, M., Lu, X., 2016. Structure-oriented substrate specificity en-
production. These methods only involve bioconversion or fermentable gineering of aldehyde-deformylating oxygenase towards aldehydes carbon chain
processes in which the whole technology generates very low CO2 length. Biotechnol. Biofuels 9, 185. https://doi.org/10.1186/s13068-016-0596-9.
Beer, L.L., Boyd, E.S., Peters, J.W., Posewitz, M.C., 2009. Engineering algae for biohy-
emission and does not produce toxic compounds. In addition to being
drogen and biofuel production. Curr. Opin. Biotechnol. 20, 264–271. https://doi.org/
green and clean for the environment, the source for hydrocarbon pro- 10.1016/j.copbio.2009.06.002.
duction by this method is also renewable. The engineered enzymes and Belcher, J., McLean, K.J., Matthews, S., Woodward, L.S., Fisher, K., Rigby, S.E., Nelson,
cells have the potential to convert abundant biomass from agriculture, D.R., Potts, D., Baynham, M.T., Parker, D.A., Leys, D., Munro, A.W., 2014. Structure
and biochemical properties of the alkene producing cytochrome P450 OleTJE
agro-industry and food waste from household for large scale production (CYP152L1) from the Jeotgalicoccus sp. 8456 bacterium. J. Biol. Chem. 289,
of alkanes and alkenes. 6535–6550. https://doi.org/10.1074/jbc.M113.527325.
Robust and efficient enzymatic cascades and metabolically en- Beld, J., Sonnenschein, E.C., Vickery, C.R., Noel, J.P., Burkart, M.D., 2014. The phos-
phopantetheinyl transferases: catalysis of a post-translational modification crucial for
gineered cells for producing hydrocarbons also have the potential to be life. Nat. Prod. Rep. 31, 61–108. https://doi.org/10.1039/c3np70054b.
revolutionary tools for sustainable and also for developing new business Bi, X., Sheng, G., Liu, X., Li, C., Fu, J., 2005. Molecular and carbon and hydrogen isotopic
models for alkane and alkene production. Rather than relying on composition of n-alkanes in plant leaf waxes. Org. Geochem. 36, 1405–1417. https://
doi.org/10.1016/j.orggeochem.2005.06.001.
countries with big oil reserves to provide crude oil and gas for hydro- Bird, S.M., Gray, J., 2003. Signals from the cuticle affect epidermal cell differentiation.
carbon production, any country with considerable biomass feedstocks New Phytol. 157. https://doi.org/10.1046/j.1469-8137.2003.00543.x.
from agriculture and agro-industries can develop a new business model Bourdenx, B., Bernard, A., Domergue, F., Pascal, S., Léger, A., Roby, D., Pervent, M., Vile,
D., Haslam, R.P., Napier, J.A., Lessire, R., Joubès, J., 2011. Overexpression of
for alkane and alkene production to be done locally. The technology Arabidopsis ECERIFERUM1 promotes wax very-long-chain alkane biosynthesis and
can even be coupled with a waste management program and can po- influences plant response to biotic and abiotic stresses. Plant Physiol. 156, 29–45.
tentially solve local waste and environmental problems. For example, https://doi.org/10.1104/pp.111.172320.
Bruder, S., Moldenhauer, E.J., Lemke, R.D., Ledesma-Amaro, R., Kabisch, J., 2019. Drop-
our research group has engaged in a project in which we propose to use
In Biofuel Production by Using Fatty Acid Photodecarboxylase from Chlorella var-
enzymatic cascades and metabolically engineered cells to increase the iabilis in the Oleaginous Yeast Yarrowia Lipolytica. bioRxiv, pp. 468876. https://doi.
value of household food waste from the community where compre- org/10.1101/468876.
hensive waste sorting and segregation is implemented (C-ROS, 2019). Buer, B.C., Paul, B., Das, D., Stuckey, J.A., Marsh, E.N., 2014. Insights into substrate and
metal binding from the crystal structure of cyanobacterial aldehyde deformylating
We are developing technology to convert food waste into hydrocarbons oxygenase with substrate bound. ACS Chem. Biol. 9, 2584–2593. https://doi.org/10.
to be used as fuels for farming engines. Such value addition to food 1021/cb500343j.
waste will support the ZeroWaste program because it will encourage the Bugalho, M., Dove, H., Kelman, W., Wood, J., Mayes, R., 2009. Plant wax alkanes and
alcohols as herbivore diet composition markers. Rangel. Ecol. Manag. 57, 259–268.
community to segregate waste so that food waste can be used for the https://doi.org/10.2111/1551-5028(2004)057[0259:PWAAAA]2.0.CO;2.
production of hydrocarbons. We believe that the bio-based method may Buschhaus, C., Herz, H., Jetter, R., 2008. Chemical composition of the epicuticular and
be relevant for industrial hydrocarbon production in the long run be- intracuticular wax layers on adaxial sides of Rosa canina leaves. Ann. Bot. 100,
1557–1564. https://doi.org/10.1093/aob/mcm255.
cause the process is sustainable, creates no environmental damage and Busta, L., Hegebarth, D., Kroc, E., Jetter, R., 2016. Changes in cuticular wax coverage and
can be used as an important tool for waste management. composition on developing Arabidopsis leaves are influenced by wax biosynthesis
In conclusion, enzymatic cascades and metabolically engineered gene expression levels and trichome density. Planta 245. https://doi.org/10.1007/
s00425-016-2603-6.
cells that can produce good yields of alkanes and alkenes are important,
C-ROS, 2019. Zero-Waste Technology for the Future. Cash Return from ZeroWaste and
clean, and sustainable technologies for the future production of fuels Segregation of Trash. https://www.c-ros.org/en/index.html.
and materials. This technology will allow countries without big oil re- Cao, Y.-X., Xiao, W.-H., Zhang, J.-L., Xie, Z., Ding, M.-Z., Yuan, Y.-J., 2016. Heterologous
biosynthesis and manipulation of alkanes in Escherichia coli. Metab. Eng. 38. https://
serves in their homeland to be capable of producing precursors for
doi.org/10.1016/j.ymben.2016.06.002.
material industries. However, more development of and research on Chatterjee, A., Hopen Eliasson, S.H., Jensen, V.R., 2018. Selective production of linear α-
new enzymes, new pathways and new types of engineered cells are olefins via catalytic deoxygenation of fatty acids and derivatives. Catal. Sci. Technol.
needed in order to obtain the whole cell biocatalysts which can provide 8, 1487–1499. https://doi.org/10.1039/C7CY02580G.
Chen, B., Lee, D.-Y., Chang, M.W., 2015. Combinatorial metabolic engineering of
greater yield with lower production costs to meet the techno-economic Saccharomyces cerevisiae for terminal alkene production. Metab. Eng. 31, 53–61.
feasibility requirement. https://doi.org/10.1016/j.ymben.2015.06.009.
Choi, Y.J., Lee, S.Y., 2013. Microbial production of short-chain alkanes. Nature 502, 571.
https://doi.org/10.1038/nature12536.
Coates, R.C., Podell, S., Korobeynikov, A., Lapidus, A., Pevzner, P., Sherman, D., Allen, E.,
Declaration of Competing Interest Gerwick, L., Gerwick, W., 2014a. Characterization of cyanobacterial hydrocarbon
composition and distribution of biosynthetic pathways. PLoS One 9, e85140. https://
doi.org/10.1371/journal.pone.0085140.
The authors declare that there is no conflict of interest. Coates, R.C., Podell, S., Korobeynikov, A., Lapidus, A., Pevzner, P., Sherman, D.H., Allen,
E.E., Gerwick, L., Gerwick, W.H., 2014b. Characterization of cyanobacterial hydro-
carbon composition and distribution of biosynthetic pathways. PLoS One 9, e85140.
https://doi.org/10.1371/journal.pone.0085140.
Acknowledgements
Coursolle, D., Lian, J., Shanklin, J., Zhao, H., 2015. Production of long chain alcohols and
alkanes upon coexpression of an acyl-ACP reductase and aldehyde-deformylating
This research was financially supported by The Thailand Research oxygenase with a bacterial type-I fatty acid synthase in E. Coli. Mol. Biosyst. 11,
2464–2472. https://doi.org/10.1039/c5mb00268k.
Fund through grants RTA5980001(to P Chaiyen), and a grant from
Das, D., Ellington, B., Paul, B., Marsh, E.N., 2014. Mechanistic insights from reaction of
Vidyasirimedhi Institute of Science and Technology (VISTEC) (to JJ, PI, alpha-oxiranyl-aldehydes with cyanobacterial aldehyde deformylating oxygenase.
WW, NA, CK, NK and PC). All figures generated with ChemBioDraw ACS Chem. Biol. 9, 570–577. https://doi.org/10.1021/cb400772q.
Ultra (Version 12.0), bioRENDER and The PyMOL Molecular graphic Davis, J.T., Moore, R.N., Imperiali, B., Pratt, A.J., Kobayashi, K., Masamune, S., Sinskey,
A.J., Walsh, C.T., Fukui, T., Tomita, K., 1987. Biosynthetic thiolase from Zoogloea
System (Version 2.3.1).

16
J. Jaroensuk, et al. Journal of Biotechnology 309 (2020) 1–19

ramigera. I. Preliminary characterization and analysis of proton transfer reaction. J. carboxylate reductases (CARs) for whole cell-mediated preparation of 3-
Biol. Chem. 262, 82–89. http://www.jbc.org/content/262/1/82.abstract. Hydroxytyrosol. ChemCatChem. https://doi.org/10.1002/cctc.201900333.
Dennig, A., Kuhn, M., Tassoti, S., Thiessenhusen, A., Gilch, S., Bülter, T., Haas, T., Hall, Hu, Q., Sommerfeld, M., Jarvis, E., Ghirardi, M., Posewitz, M., Seibert, M., Darzins, A.,
M., Faber, K., 2015. Oxidative decarboxylation of short-chain fatty acids to 1- 2008. Microalgal triacylglycerols as feedstocks for biofuel production: perspectives
Alkenes. Angew. Chem. Int. Ed. 54, 8819–8822. https://doi.org/10.1002/anie. and advances. Plant J. 54, 621–639. https://doi.org/10.1111/j.1365-313X.2008.
201502925. 03492.x.
Dhameja, S., 2002. Electric vehicle Battery fast charging. In: Dhameja, S. (Ed.), Electric Huijbers, M.M.E., Zhang, W., Tonin, F., Hollmann, F., 2018. Light-driven enzymatic
Vehicle Battery Systems. Newnes, Woburn, pp. 95–114. decarboxylation of fatty acids. Angew. Chem. Int. Ed. 57, 13648–13651. https://doi.
Dodd, R.S., Rafii, Z.A., Power, A.B., 1998. Ecotypic adaptation in Austrocedrus chilensis in org/10.1002/anie.201807119.
cuticular hydrocarbon composition. New Phytol. 138, 699–708. https://doi.org/10. Jaroensuk, J., Intasian, P., Kiattisewee, C., Munkajohnpon, P., Chunthaboon, P.,
1046/j.1469-8137.1998.00142.x. Buttranon, S., Trisrivirat, D., Wongnate, T., Maenpuen, S., Tinikul, R., Chaiyen, P.,
Dominguez, E., Cuartero, J., Heredia, A., 2011. An overview on plant cuticle bio- 2019. Addition of formate dehydrogenase increases the production of renewable
mechanics. Plant Sci. 181, 77–84. https://doi.org/10.1016/j.plantsci.2011.04.016. alkane from an engineered metabolic pathway. J. Biol. Chem. 294, 11536–11548.
Dragota, S., Riederer, M., 2007. Epicuticular wax crystals of Wollemia nobilis: morphology https://doi.org/10.1074/jbc.RA119.008246.
and chemical composition. Ann. Bot. 100, 225–231. https://doi.org/10.1093/aob/ Jetter, R., Kunst, L., 2008. Plant surface lipid biosynthetic pathways and their utility for
mcm120. metabolic engineering of waxes and hydrocarbon biofuels. Plant J. 54, 670–683.
Evans, A., Strezov, V., Evans, T.J., 2009. Assessment of sustainability indicators for re- https://doi.org/10.1111/j.1365-313X.2008.03467.x.
newable energy technologies. Renew. Sustain. Energy Rev. 13, 1082–1088. https:// Jetter, R., Riederer, M., 2011. Cuticular waxes from the leaves and fruit capsules of eight
doi.org/10.1016/j.rser.2008.03.008. Papaveraceae species. Can. J. Bot. 74, 419–430. https://doi.org/10.1139/b96-052.
Diefendorf, A.F., Freeman, K., Wing, S., Graham, H., 2011. Production of n-alkyl lipids in Jia, C., Li, M., Li, J., Zhang, J., Zhang, H., Cao, P., Pan, X., Lu, X., Chang, W., 2015.
living plants and implications for the geologic past. Geochimica Et Cosmochimica Structural insights into the catalytic mechanism of aldehyde-deformylating oxyge-
Acta - Geochim Cosmochim Acta 75, 7472–7485. https://doi.org/10.1016/j.gca. nases. Protein Cell 6, 55–67. https://doi.org/10.1007/s13238-014-0108-2.
2011.09.028. Jimenez-Diaz, L., Caballero, A., Perez-Hernandez, N., Segura, A., 2017. Microbial alkane
Fang, B., Xu, H., Liu, Y., Qi, F., Zhang, W., Chen, H., Wang, C., Wang, Y., Yang, W., Li, S., production for jet fuel industry: motivation, state of the art and perspectives. Microb.
2017. Mutagenesis and redox partners analysis of the P450 fatty acid decarboxylase Biotechnol. 10, 103–124. https://doi.org/10.1111/1751-7915.12423.
OleTJE. Sci. Rep. 7, 44258. https://doi.org/10.1038/srep44258. Kallio, P., Pasztor, A., Thiel, K., Akhtar, M.K., Jones, P.R., 2014. An engineered pathway
Fatma, Z., Hartman, H., Poolman, M.G., Fell, D.A., Srivastava, S., Shakeel, T., Yazdani, for the biosynthesis of renewable propane. Nat. Commun. 5, 4731. https://doi.org/
S.S., 2018. Model-assisted metabolic engineering of Escherichia coli for long chain 10.1038/ncomms5731.
alkane and alcohol production. Metab. Eng. 46, 1–12. https://doi.org/10.1016/j. Kang, M.-K., Nielsen, J., 2017. Biobased production of alkanes and alkenes through
ymben.2018.01.002. metabolic engineering of microorganisms. J. Ind. Microbiol. Biotechnol. 44, 613–622.
Finnigan, W., Thomas, A., Cromar, H., Gough, B., Snajdrova, R., Adams, J.P., Littlechild, https://doi.org/10.1007/s10295-016-1814-y.
J.A., Harmer, N.J., 2017. Characterization of carboxylic acid reductases as enzymes Khara, B., Menon, N., Levy, C., Mansell, D., Das, D., Marsh, E.N., Leys, D., Scrutton, N.S.,
in the toolbox for synthetic chemistry. ChemCatChem 9, 1005–1017. https://doi.org/ 2013. Production of propane and other short-chain alkanes by structure-based en-
10.1002/cctc.201601249. gineering of ligand specificity in aldehyde-deformylating oxygenase. Chembiochem
Foo, J.L., Susanto, A.V., Keasling, J.D., Leong, S.S., Chang, M.W., 2017. Whole-cell bio- 14, 1204–1208. https://doi.org/10.1002/cbic.201300307.
catalytic and de novo production of alkanes from free fatty acids in Saccharomyces Kim, H.M., Chae, T.U., Choi, S.Y., Kim, W.J., Lee, S.Y., 2019. Engineering of an oleagi-
cerevisiae. Biotechnol. Bioeng. 114, 232–237. https://doi.org/10.1002/bit.25920. nous bacterium for the production of fatty acids and fuels. Nat. Chem. Biol. 15,
Frias, J.A., Goblirsch, B.R., Wackett, L.P., Wilmot, C.M., 2010. Cloning, purification, 721–729. https://doi.org/10.1038/s41589-019-0295-5.
crystallization and preliminary X-ray diffraction of the OleC protein from Kohli, K., Prajapati, R., Sharma, B.K., 2019. Bio-based chemicals from renewable biomass
Stenotrophomonas maltophilia involved in head-to-head hydrocarbon biosynthesis. for integrated biorefineries. Energies 12, 233. https://doi.org/10.3390/en12020233.
Acta Crystallogr. Sect. F Struct. Biol. Cryst. Commun. 66, 1108–1110. https://doi. Kramer, L., Hankore, E.D., Liu, Y., Liu, K., Jimenez, E., Guo, J., Niu, W., 2018.
org/10.1107/S1744309110031751. Characterization of carboxylic acid reductases for biocatalytic synthesis of industrial
Frias, J.A., Richman, J.E., Erickson, J.S., Wackett, L.P., 2011. Purification and char- chemicals. Chembiochem 19, 1452–1460. https://doi.org/10.1002/cbic.201800157.
acterization of OleA from Xanthomonas campestris and demonstration of a non-dec- Kudo, H., Nawa, R., Hayashi, Y., Arai, M., 2016. Comparison of aldehyde-producing ac-
arboxylative claisen condensation reaction. J. Biol. Chem. 286, 10930–10938. tivities of cyanobacterial acyl-(acyl carrier protein) reductases. Biotechnol. Biofuels
https://doi.org/10.1074/jbc.M110.216127. 9, 234. https://doi.org/10.1186/s13068-016-0644-5.
Gahloth, D., Dunstan, M.S., Quaglia, D., Klumbys, E., Lockhart-Cairns, M.P., Hill, A.M., Lee, D., Yamada, A., Sugimoto, H., Matsunaga, I., Ogura, H., Ichihara, K., Adachi, S.-I.,
Derrington, S.R., Scrutton, N.S., Turner, N.J., Leys, D., 2017. Structures of carboxylic Park, S.-Y., Shiro, Y., 2003. Substrate recognition and molecular mechanism of fatty
acid reductase reveal domain dynamics underlying catalysis. Nat. Chem. Biol. 13, acid hydroxylation by cytochrome P450 from Bacillus subtilis - Crystallographic,
975–981. https://doi.org/10.1038/nchembio.2434. spectroscopic, and mutational studies. J. Biol. Chem. 278, 9761–9767. https://doi.
Gao, L., Huang, Y., 2013. Inverse gradients in leaf wax δD and δ13C values along grass org/10.1074/jbc.M211575200.
blades of Miscanthus sinensis: Implications for leaf wax reproduction and plant phy- Lehtinen, T., Efimova, E., Santala, S., Santala, V., 2018. Improved fatty aldehyde and wax
siology. Oecologia 172. https://doi.org/10.1007/s00442-012-2506-6. ester production by overexpression of fatty acyl-CoA reductases. Microb. Cell Fact.
Ghanta, M., Fahey, D., Subramaniam, B., 2014. Environmental impacts of ethylene pro- 17, 19. https://doi.org/10.1186/s12934-018-0869-z.
duction from diverse feedstocks and energy sources. Appl. Petrochem. Res. 4, León-Bañares, R., González-Ballester, D., Galván, A., Fernández, E., 2004. Transgenic
167–179. https://doi.org/10.1007/s13203-013-0029-7. microalgae as green cell-factories. Trends Biotechnol. 22, 45–52. https://doi.org/10.
Grant, J.L., Hsieh, C.H., Makris, T.M., 2015. Decarboxylation of fatty acids to terminal 1016/j.tibtech.2003.11.003.
alkenes by cytochrome P450 compound I. J. Am. Chem. Soc. 137, 4940–4943. Li, F., Yang, K., Xu, Y., Qiao, Y., Yan, Y., Yan, J., 2019. A genetically-encoded synthetic
https://doi.org/10.1021/jacs.5b01965. self-assembled multienzyme complex of lipase and P450 fatty acid decarboxylase for
Grim, R.G., To, A.T., Farberow, C.A., Hensley, J.E., Ruddy, D.A., Schaidle, J.A., 2019. efficient bioproduction of fatty alkenes. Bioresour. Technol. 272, 451–457. https://
Growing the bioeconomy through catalysis: a review of recent advancements in the doi.org/10.1016/j.biortech.2018.10.067.
production of fuels and chemicals from syngas-derived oxygenates. ACS Catal. 9, Li, N., Chang, W.-C., Warui, D.J., Booker, S., Krebs, C., Bollinger, J., 2012. Evidence for
4145–4172. https://doi.org/10.1021/acscatal.8b03945. only oxygenative cleavage of aldehydes to alk(a/e)nes and formate by cyanobacterial
Gu, L., Wang, B., Kulkarni, A., Gehret, J.J., Lloyd, K.R., Gerwick, L., Gerwick, W.H., Wipf, aldehyde decarbonylases. Biochemistry 51, 7908–7916. https://doi.org/10.1021/
P., Håkansson, K., Smith, J.L., Sherman, D.H., 2009. Polyketide decarboxylative bi300912n.
chain termination preceded by O-Sulfonation in curacin a biosynthesis. J. Am. Chem. Li, N., Nørgaard, H., Warui, D.M., Booker, S.J., Krebs, C., Bollinger, J.M., 2011.
Soc. 131, 16033–16035. https://doi.org/10.1021/ja9071578. Conversion of fatty aldehydes to Alka(e)nes and formate by a cyanobacterial alde-
Guhling, O., Kinzler, C., Dreyer, M., Bringmann, G., Jetter, R., 2005. Surface composition hyde decarbonylase: cryptic redox by an unusual dimetal oxygenase. J. Am. Chem.
of myrmecophilic plants: cuticular wax and glandular trichomes on leaves of Macaranga Soc. 133, 6158–6161. https://doi.org/10.1021/ja2013517.
tanarius. J. Chem. Ecol. 31, 2323–2341. https://doi.org/10.1007/s10886-005- Lin, F., Das, D., Lin, X.N., Marsh, E.N., 2013. Aldehyde-forming fatty acyl-CoA reductase
7104-1. from cyanobacteria: expression, purification and characterization of the recombinant
Harwood, J.L., Guschina, I.A., 2009. The versatility of algae and their lipid metabolism. enzyme. FEBS J. 280, 4773–4781. https://doi.org/10.1111/febs.12443.
Biochimie 91, 679–684. https://doi.org/10.1016/j.biochi.2008.11.004. Liu, B., Benning, C., 2013. Lipid metabolism in microalgae distinguishes itself. Curr. Opin.
Hegebarth, D., Buschhaus, C., Joubès, J., Thoraval, D., Jetter, R., 2017. Arabidopsis ke- Biotechnol. 24, 300–309. https://doi.org/10.1016/j.copbio.2012.08.008.
toacyl-CoA synthase 16 forms C36/C38 acyl precursors for leaf trichome and pave- Liu, Y.-Y., Chi, Z., Wang, Z.-P., Liu, G.-L., Chi, Z.-M., 2014a. Heavy oils, principally long-
ment surface wax. Plant Cell Environ. 40. https://doi.org/10.1111/pce.12981. chain n-alkanes secreted by Aureobasidium pullulans var. melanogenum strain P5 iso-
Hegebarth, D., Jetter, R., 2017. Cuticular waxes of Arabidopsis thaliana shoots: cell-type- lated from mangrove system. J. Ind. Microbiol. Biotechnol. 41, 1329–1337. https://
Specific composition and biosynthesis. Plants (Basel, Switzerland) 6, 27. https://doi. doi.org/10.1007/s10295-014-1484-6.
org/10.3390/plants6030027. Liu, Y., Wang, C., Yan, J., Zhang, W., Guan, W., Lu, X., Li, S., 2014b. Hydrogen peroxide-
Herman, N.A., Zhang, W., 2016. Enzymes for fatty acid-based hydrocarbon biosynthesis. independent production of α-alkenes by OleTJE P450 fatty acid decarboxylase.
Curr. Opin. Chem. Biol. 35, 22–28. https://doi.org/10.1016/j.cbpa.2016.08.009. Biotechnol. Biofuels 7, 28. https://doi.org/10.1186/1754-6834-7-28.
Hoeven, R., Hughes, J.M.X., Amer, M., Wojcik, E.Z., Tait, S., Faulkner, M., Yunus, I.S., Lockey, K.H., 1988. Lipids of the insect cuticle: origin, composition and function. Comp.
Hardman, S.J.O., Johannissen, L.O., Chen, G.-Q., Smith, M.H., Jones, P.R., Toogood, Biochem. Physiol. Part B Comp. Biochem. 89, 595–645. https://doi.org/10.1016/
H.S., Scrutton, N.S., 2019. Distributed Biomanufacturing of Liquefied Petroleum Gas. 0305-0491(88)90305-7.
bioRxiv, pp. 640474. https://doi.org/10.1101/640474. Luo, J., Lehtinen, T., Efimova, E., Santala, V., Santala, S., 2019. Synthetic metabolic
Horvat, M., Fritsche, S., Kourist, R., Winkler, M., 2019. Characterization of type IV pathway for the production of 1-alkenes from lignin-derived molecules. Microb. Cell

17
J. Jaroensuk, et al. Journal of Biotechnology 309 (2020) 1–19

Fact. 18, 48. https://doi.org/10.1186/s12934-019-1097-x. doi.org/10.1021/acscatal.5b01842.


MacLean, M., Nadeau, J., Gurnea, T., Tittiger, C., Blomquist, G.J., 2018. Mountain pine Rui, Z., Li, X., Zhu, X., Liu, J., Domigan, B., Barr, I., Cate, J.H.D., Zhang, W., 2014.
beetle (Dendroctonus ponderosae) CYP4Gs convert long and short chain alcohols and Microbial biosynthesis of medium-chain 1-alkenes by a nonheme iron oxidase. Proc.
aldehydes to hydrocarbons. Insect Biochem. Mol. Biol. 102, 11–20. https://doi.org/ Natl. Acad. Sci. 111, 18237. https://doi.org/10.1073/pnas.1419701112.
10.1016/j.ibmb.2018.09.005. Samuels, A., Kunst, L., Jetter, R., 2008. Sealing plant surfaces: cuticular wax formation by
Malcata, F.X., 2011. Microalgae and biofuels: A promising partnership? Trends epidermal cells. Annu. Rev. Plant Biol. 59, 683–707. https://doi.org/10.1146/
Biotechnol. 29, 542–549. https://doi.org/10.1016/j.tibtech.2011.05.005. annurev.arplant.59.103006.093219.
Matsunaga, I., Sumimoto, T., Ueda, A., Kusunose, E., Ichihara, K., 2000. Fatty acid-spe- Santala, S., Efimova, E., Kivinen, V., Larjo, A., Aho, T., Karp, M., Santala, V., 2011.
cific, regiospecific, and stereospecific hydroxylation by cytochrome P450 Improved triacylglycerol production in Acinetobacter baylyi ADP1 by metabolic en-
(CYP152B1) from Sphingomonas paucimobilis: substrate structure required for alpha- gineering. Microb. Cell Fact. 10, 36. https://doi.org/10.1186/1475-2859-10-36.
hydroxylation. Lipids 35, 365–371. https://doi.org/10.1007/s11745-000-533-y. Santala, S., Efimova, E., Koskinen, P., Karp, M.T., Santala, V., 2014. Rewiring the wax
Matthews, S., Tee, K.L., Rattray, N.J., McLean, K.J., Leys, D., Parker, D.A., Blankley, R.T., ester production pathway of Acinetobacter baylyi ADP1. ACS Synth. Biol. 3, 145–151.
Munro, A.W., 2017. Production of alkenes and novel secondary products by P450 https://doi.org/10.1021/sb4000788.
OleTJE using novel H2O2-generating fusion protein systems. FEBS Lett. 591, Schirmer, A., Rude, M.A., Li, X., Popova, E., del Cardayre, S.B., 2010. Microbial bio-
737–750. https://doi.org/10.1002/1873-3468.12581. synthesis of alkanes. Science 329, 559. https://doi.org/10.1126/science.1187936.
Mendez-Perez, D., Begemann, M.B., Pfleger, B.F., 2011. Modular synthase-encoding gene Schirmer, W.A., Rude, A.M., Shane, B.A., 2014. In: Patent, U.S. (Ed.), Methods and
involved in α-olefin biosynthesis in Synechococcus sp. Strain PCC 7002. Appl. Compositions for Producing Alkanes and Alkenes. REG Life Science, LLC., United
Environ. Microbiol. 77, 4264–4267. https://doi.org/10.1128/AEM.00467-11. States.
Menon, N., Pasztor, A., Menon, B.R., Kallio, P., Fisher, K., Akhtar, M.K., Leys, D., Jones, Schwendenwein, D., Fiume, G., Weber, H., Rudroff, F., Winkler, M., 2016. Selective en-
P.R., Scrutton, N.S., 2015. A microbial platform for renewable propane synthesis zymatic transformation to aldehydes in vivo by fungal carboxylate reductase from
based on a fermentative butanol pathway. Biotechnol. Biofuels 8, 61. https://doi.org/ Neurospora crassa. Adv. Synth. Catal. 358, 3414–3421. https://doi.org/10.1002/
10.1186/s13068-015-0231-1. adsc.201600914.
Mihailova, A., Abbado, D., Pedentchouk, N., 2015. Differences in n-alkane profiles be- Schwendenwein, D., Ressmann, A.K., Doerr, M., Höhne, M., Bornscheuer, U.T.,
tween olives and olive leaves as potential indicators for the assessment of olive leaf Mihovilovic, M.D., Rudroff, F., Winkler, M., 2019. Random mutagenesis-driven im-
presence in virgin olive oils. Eur. J. Lipid Sci. Technol. 117. https://doi.org/10.1002/ provement of carboxylate reductase activity using an amino benzamidoxime-medi-
ejlt.201400406. ated high-throughput assay. Adv. Synth. Catal. 361, 2544–2549. https://doi.org/10.
Moss, N.A., Bertin, M.J., Kleigrewe, K., Leão, T.F., Gerwick, L., Gerwick, W.H., 2016. 1002/adsc.201900155.
Integrating mass spectrometry and genomics for cyanobacterial metabolite discovery. Scrutton, N.S., 2017. Enzymes make light work of hydrocarbon production. Science 357,
J. Ind. Microbiol. Biotechnol. 43, 313–324. https://doi.org/10.1007/s10295-015- 872–873. https://doi.org/10.1126/science.aao4399.
1705-7. Sheppard, M.J., Kunjapur, A.M., Prather, K.L.J., 2016. Modular and selective biosynthesis
Mpuru, S., Reed, J.R., Reitz, R.C., Blomquist, G.J., 1996. Mechanism of hydrocarbon of gasoline-range alkanes. Metab. Eng. 33, 28–40. https://doi.org/10.1016/j.ymben.
biosynthesis from aldehyde in selected insect species: requirement for O2 and NADPH 2015.10.010.
and carbonyl group released as CO2. Insect Biochem. Mol. Biol. 26, 203–208. https:// Shoji, O., Fujishiro, T., Nakajima, H., Kim, M., Nagano, S., Shiro, Y., Watanabe, Y., 2007.
doi.org/10.1016/0965-1748(95)00086-0. Hydrogen peroxide dependent monooxygenations by tricking the substrate recogni-
Munro, A.W., McLean, K.J., Grant, J.L., Makris, T.M., 2018. Structure and function of the tion of cytochrome P450BSbeta. Angew. Chem. Int. Ed. Engl. 46, 3656–3659. https://
cytochrome P450 peroxygenase enzymes. Biochem. Soc. Trans. 46, 183–196. https:// doi.org/10.1002/anie.200700068.
doi.org/10.1042/BST20170218. Sivaramakrishnan, R., Incharoensakdi, A., 2018. Microalgae as feedstock for biodiesel
Nadiminti, P., Rookes, J., Boyd, B., Cahill, D., 2015. Confocal laser scanning microscopy production under ultrasound treatment - A review. Bioresour. Technol. 250,
elucidation of the micromorphology of the leaf cuticle and analysis of its chemical 877–887. https://doi.org/10.1016/j.biortech.2017.11.095.
composition. Protoplasma 252. https://doi.org/10.1007/s00709-015-0777-6. Sligar, S.G., Gunsalus, I.C., 1979. Proton coupling in the cytochrome P-450 spin and redox
Ng, T.K., Yu, A.Q., Ling, H., Pratomo Juwono, N.K., Choi, W.J., Leong, S.S.J., Chang, equilibria. Biochemistry 18, 2290–2295. https://doi.org/10.1021/bi00578a024.
M.W., 2019. Engineering Yarrowia lipolytica towards food waste bioremediation: Sorigue, D., Legeret, B., Cuine, S., Blangy, S., Moulin, S., Billon, E., Richaud, P., Brugiere,
production of fatty acid ethyl esters from vegetable cooking oil. J. Biosci. Bioeng. S., Coute, Y., Nurizzo, D., Muller, P., Brettel, K., Pignol, D., Arnoux, P., Li-Beisson, Y.,
https://doi.org/10.1016/j.jbiosc.2019.06.009. Peltier, G., Beisson, F., 2017. An algal photoenzyme converts fatty acids to hydro-
Patrikainen, P., Carbonell, V., Thiel, K., Aro, E.M., Kallio, P., 2017. Comparison of or- carbons. Science 357, 903–907. https://doi.org/10.1126/science.aan6349.
thologous cyanobacterial aldehyde deformylating oxygenases in the production of Sorigué, D., Légeret, B., Cuiné, S., Morales, P., Mirabella, B., Guédeney, G., Li-Beisson, Y.,
volatile C3-C7 alkanes in engineered E. coli. Metab. Eng. Commun. 5, 9–18. https:// Jetter, R., Peltier, G., Beisson, F., 2016. Microalgae synthesize hydrocarbons from
doi.org/10.1016/j.meteno.2017.05.001. long-chain fatty acids via a light-dependent pathway. Plant Physiol. 171, 2393–2405.
Paul, B., Das, D., Ellington, B., Marsh, E.N., 2013. Probing the mechanism of cyano- https://doi.org/10.1104/pp.16.00462.
bacterial aldehyde decarbonylase using a cyclopropyl aldehyde. J. Am. Chem. Soc. Stolterfoht, H., Schwendenwein, D., Sensen, C.W., Rudroff, F., Winkler, M., 2017. Four
135, 5234–5237. https://doi.org/10.1021/ja3115949. distinct types of E.C. 1.2.1.30 enzymes can catalyze the reduction of carboxylic acids
Qiu, Y., Tittiger, C., Wicker-Thomas, C., Le Goff, G., Young, S., Wajnberg, E., Fricaux, T., to aldehydes. J. Biotechnol. 257, 222–232. https://doi.org/10.1016/j.jbiotec.2017.
Taquet, N., Blomquist, G.J., Feyereisen, R., 2012. An insect-specific P450 oxidative 02.014.
decarbonylase for cuticular hydrocarbon biosynthesis. Proc. Natl. Acad. Sci. U. S. A. Stolterfoht, H., Steinkellner, G., Schwendenwein, D., Pavkov-Keller, T., Gruber, K.,
109, 14858–14863. https://doi.org/10.1073/pnas.1208650109. Winkler, M., 2018. Identification of key residues for enzymatic carboxylate reduc-
Qu, G., Fu, M., Zhao, L., Liu, B., Liu, P., Fan, W., Ma, J.A., Sun, Z., 2019. Computational tion. Front. Microbiol. 9, 250. https://doi.org/10.3389/fmicb.2018.00250.
insights into the catalytic mechanism of bacterial carboxylic acid reductase. J. Chem. Sukovich, D.J., Seffernick, J.L., Richman, J.E., Hunt, K.A., Gralnick, J.A., Wackett, L.P.,
Inf. Model. 59, 832–841. https://doi.org/10.1021/acs.jcim.8b00763. 2010. Structure, function, and insights into the biosynthesis of a head-to-Head hy-
Qu, G., Guo, J., Yang, D., Sun, Z., 2018. Biocatalysis of carboxylic acid reductases: phy- drocarbon in Shewanella oneidensis strain MR-1. Appl. Environ. Microbiol. 76, 3842.
logenesis, catalytic mechanism and potential applications. Green Chem. 20, 777–792. https://doi.org/10.1128/AEM.00433-10.
https://doi.org/10.1039/C7GC03046K. Thangaraj, B., Solomon, P.R., Muniyandi, B., Ranganathan, S., Lin, L., 2018. Catalysis in
Rajakovich, L.J., Norgaard, H., Warui, D.M., Chang, W.C., Li, N., Booker, S.J., Krebs, C., biodiesel production—a review. Clean Energy 3, 2–23. https://doi.org/10.1093/ce/
Bollinger Jr., J.M., Pandelia, M.E., 2015. Rapid reduction of the diferric-perox- zky020.
yhemiacetal intermediate in aldehyde-deformylating oxygenase by a cyanobacterial Thome, T., Erhardt, M.C., Leme-Kraus, A., Bedran-Russo, A.K.B., Al‐Bakri, I., Bertassoni,
ferredoxin: evidence for a free-radical mechanism. J. Am. Chem. Soc. 137, L., 2015. Emerging polymers in dentistry. Adv. Polym. Med. 265–295. https://doi.
11695–11709. https://doi.org/10.1021/jacs.5b06345. org/10.1007/978-3-319-12478-0_9.
Ramseur, J.L., Leggett, J.A., 2019. Attaching a Price to Greenhouse Gas Emissions with a Titah, H.S., Abdullah, S.R.S., Mushrifah, I., Anuar, N., Basri, H., Mukhlisin, M., 2013.
Carbon Tax or Emissions Fee: Considerations and Potential Impacts. Congressional Effect of applying rhizobacteria and fertilizer on the growth of Ludwigia octovalvis for
Research Service. https://www.hsdl.org/?view&did=823346. arsenic uptake and accumulation in phytoremediation. Ecol. Eng. 58, 303–313.
Reed, J.R., Vanderwel, D., Choi, S., Pomonis, J.G., Reitz, R.C., Blomquist, G.J., 1994. https://doi.org/10.1016/j.ecoleng.2013.07.018.
Unusual mechanism of hydrocarbon formation in the housefly: cytochrome P450 Toogood, H.S., Scrutton, N.S., 2018. Retooling microorganisms for the fermentative
converts aldehyde to the sex pheromone component (Z)-9-tricosene and CO2. Proc. production of alcohols. Curr. Opin. Biotechnol. 50, 1–10. https://doi.org/10.1016/j.
Natl. Acad. Sci. U. S. A. 91, 10000–10004. https://doi.org/10.1073/pnas.91.21. copbio.2017.08.010.
10000. Tsiropoulos, I., Tarvydas, D., Lebedeva, N., 2018. Li-ion Batteries for Mobility and
Reiser, S., Somerville, C., 1997. Isolation of mutants of Acinetobacter calcoaceticus defi- Stationary Storage Applications Scenarios for Costs and Market Growth. https://doi.
cient in wax ester synthesis and complementation of one mutation with a gene en- org/10.2760/87175.
coding a fatty acyl coenzyme A reductase. J. Bacteriol. 179, 2969–2975. https://doi. Wada, H., Murata, N., 1998. Membrane lipids in cyanobacteria. In: Paul-André, S., Norio,
org/10.1128/jb.179.9.2969-2975.1997. M. (Eds.), Lipids in Photosynthesis: Structure, Function and Genetics. Springer
Rogner, H., 1997. An assessment of world hydrocarbon resources. Ann. Rev. Energy Netherlands, Dordrecht, pp. 65–81. https://doi.org/10.1007/0-306-48087-5_4.
Environ. 22. https://doi.org/10.1146/annurev.energy.22.1.217. Wang, J., Yu, H., Song, X., Zhu, K., 2018. The influence of fatty acid supply and aldehyde
Rude, M.A., Baron, T.S., Brubaker, S., Alibhai, M., Del Cardayre, S.B., Schirmer, A., 2011. reductase deletion on cyanobacteria alkane generating pathway in Escherichia coli. J.
Terminal olefin (1-alkene) biosynthesis by a novel p450 fatty acid decarboxylase Ind. Microbiol. Biotechnol. 45, 329–334. https://doi.org/10.1007/s10295-018-
from Jeotgalicoccus species. Appl. Environ. Microbiol. 77, 1718–1727. https://doi. 2032-6.
org/10.1128/AEM.02580-10. Wang, Q., Bao, L., Jia, C., Li, M., Li, J.J., Lu, X., 2017. Identification of residues important
Rui, Z., Harris, N.C., Zhu, X., Huang, W., Zhang, W., 2015. Discovery of a family of de- for the activity of aldehyde-deformylating oxygenase through investigation into the
saturase-like enzymes for 1-Alkene biosynthesis. ACS Catal. 5, 7091–7094. https:// structure-activity relationship. BMC Biotechnol. 17, 31. https://doi.org/10.1186/

18
J. Jaroensuk, et al. Journal of Biotechnology 309 (2020) 1–19

s12896-017-0351-8. Photobiocatalytic decarboxylation for olefin synthesis. Chem. Commun. (Camb.) 51,
Wang, W., Liu, X., Lu, X., 2013. Engineering cyanobacteria to improve photosynthetic 1918–1921. https://doi.org/10.1039/c4cc07276f.
production of alka(e)nes. Biotechnol. Biofuels 6, 69. https://doi.org/10.1186/1754- Zhang, J., Lu, X., Li, J.-J., 2013. Conversion of fatty aldehydes into alk (a/e)nes by in
6834-6-69. vitroreconstituted cyanobacterial aldehyde-deformylating oxygenase with the cog-
Warui, D.M., Pandelia, M.E., Rajakovich, L.J., Krebs, C., Bollinger Jr, J.M., Booker, S.J., nate electron transfer system. Biotechnol. Biofuels 6, 86. https://doi.org/10.1186/
2015. Efficient delivery of long-chain fatty aldehydes from the Nostoc punctiforme 1754-6834-6-86.
acyl-acyl carrier protein reductase to its cognate aldehyde-deformylating oxygenase. Zhang, W., Ma, M., Huijbers, M.M.E., Filonenko, G.A., Pidko, E.A., van Schie, M., de Boer,
Biochemistry 54, 1006–1015. https://doi.org/10.1021/bi500847u. S., Burek, B.O., Bloh, J.Z., van Berkel, W.J.H., Smith, W.A., Hollmann, F., 2019.
Waugh, M.W., Marsh, E.N., 2014. Solvent isotope effects on alkane formation by cya- Hydrocarbon synthesis via photoenzymatic decarboxylation of carboxylic acids. J.
nobacterial aldehyde deformylating oxygenase and their mechanistic implications. Am. Chem. Soc. 141, 3116–3120. https://doi.org/10.1021/jacs.8b12282.
Biochemistry 53, 5537–5543. https://doi.org/10.1021/bi5005766. Zhang, Y., Togamura, Y., Otsuki, K., 2004. Study on the n-alkane patterns in some grasses
Wijffels, R.H., Kruse, O., Hellingwerf, K.J., 2013. Potential of industrial biotechnology and factors affecting the n-alkane patterns. J. Agric. Sci. 142, 469–475. https://doi.
with cyanobacteria and eukaryotic microalgae. Curr. Opin. Biotechnol. 24, 405–413. org/10.1017/S0021859604004526.
https://doi.org/10.1016/j.copbio.2013.04.004. Zheng, H., Rowland, O., Kunst, L., 2005. Disruptions of the Arabidopsis Enoyl-CoA re-
Winkler, M., 2018. Carboxylic acid reductase enzymes (CARs). Curr. Opin. Chem. Biol. ductase gene reveal an essential role for very-long-chain fatty acid synthesis in cell
43, 23–29. https://doi.org/10.1016/j.cbpa.2017.10.006. expansion during plant morphogenesis. Plant Cell 17, 1467–1481. https://doi.org/
Wood, A.J.L., Weise, N.J., Frampton, J.D., Dunstan, M.S., Hollas, M.A., Derrington, S.R., 10.1105/tpc.104.030155.
Lloyd, R.C., Quaglia, D., Parmeggiani, F., Leys, D., Turner, N.J., Flitsch, S.L., 2017. Zhou, Y.J., Buijs, N.A., Zhu, Z., Gomez, D.O., Boonsombuti, A., Siewers, V., Nielsen, J.,
Adenylation activity of carboxylic acid reductases enables the synthesis of Amides. 2016a. Harnessing yeast peroxisomes for biosynthesis of fatty-acid-Derived biofuels
Angew. Chem. Int. Ed. Engl. 56, 14498–14501. https://doi.org/10.1002/anie. and chemicals with relieved side-pathway competition. J. Am. Chem. Soc. 138,
201707918. 15368–15377. https://doi.org/10.1021/jacs.6b07394.
Xin, F.-H., Zhang, Y., Xue, S.-J., Chi, Z., Liu, G.-L., Hu, Z., Chi, Z.-M., 2017. Heavy oils Zhou, Y.J., Buijs, N.A., Zhu, Z., Qin, J., Siewers, V., Nielsen, J., 2016b. Production of fatty
(mainly alkanes) over-production from inulin by Aureobasidium melanogenum 9-1 and acid-derived oleochemicals and biofuels by synthetic yeast cell factories. Nat.
its transformant 88 carrying an inulinase gene. Renew. Energy 105, 561–568. Commun. 7, 11709. https://doi.org/10.1038/ncomms11709.
https://doi.org/10.1016/j.renene.2017.01.004. Zhou, Y.J., Hu, Y., Zhu, Z., Siewers, V., Nielsen, J., 2018a. Engineering 1-Alkene bio-
Xu, J., Hu, Y., Fan, J., Arkin, M., Li, D., Peng, Y., Xu, W., Lin, X., Wu, Q., 2019. Light- synthesis and secretion by dynamic regulation in yeast. ACS Synth. Biol. 7, 584–590.
driven kinetic resolution of alpha-functionalized carboxylic acids enabled by an en- https://doi.org/10.1021/acssynbio.7b00338.
gineered fatty acid photodecarboxylase. Angew. Chem. Int. Ed. Engl. 58, 8474–8478. Zhou, Y.J., Kerkhoven, E.J., Nielsen, J., 2018b. Barriers and opportunities in bio-based
https://doi.org/10.1002/anie.201903165. production of hydrocarbons. Nat. Energy 3, 925–935. https://doi.org/10.1038/
Yan, J., Liu, Y., Wang, C., Han, B., Li, S., 2015. Assembly of lipase and P450 fatty acid s41560-018-0197-x.
decarboxylase to constitute a novel biosynthetic pathway for production of 1-alkenes Zhu, T., Scalvenzi, T., Sassoon, N., Lu, X., Gugger, M., 2018. Terminal olefin profiles and
from renewable triacylglycerols and oils. Biotechnol. Biofuels 8, 34. https://doi.org/ phylogenetic analyses of olefin synthases of diverse cyanobacterial species. Appl.
10.1186/s13068-015-0219-x. Environ. Microbiol. 84. https://doi.org/10.1128/AEM.00425-18.
Yu, H., Kim, J., Lee, C., 2019. Nutrient removal and microalgal biomass production from Zhu, Z., Zhou, Y.J., Kang, M.K., Krivoruchko, A., Buijs, N.A., Nielsen, J., 2017. Enabling
different anaerobic digestion effluents with Chlorella species. Sci. Rep. 9, 6123. the synthesis of medium chain alkanes and 1-alkenes in yeast. Metab. Eng. 44, 81–88.
https://doi.org/10.1038/s41598-019-42521-2. https://doi.org/10.1016/j.ymben.2017.09.007.
Zachos, I., Gassmeyer, S.K., Bauer, D., Sieber, V., Hollmann, F., Kourist, R., 2015.

19

You might also like