You are on page 1of 18

Molecular Plant

Review Article

Vanillin–Bioconversion and Bioengineering of the


Most Popular Plant Flavor and Its De Novo
Biosynthesis in the Vanilla Orchid
Nethaji J. Gallage1,2,3 and Birger Lindberg Møller1,2,3,4,*
1
VILLUM Research Center for Plant Plasticity, Department of Plant and Environmental Sciences, University of Copenhagen, 40 Thorvaldsensvej, DK-1871
Frederiksberg C, Copenhagen, Denmark
2
Center for Synthetic Biology ‘‘bioSYNergy’’, Department of Plant and Environmental Sciences, University of Copenhagen, 40 Thorvaldsensvej, DK-1871
Frederiksberg C, Copenhagen, Denmark
3
Plant Biochemistry Laboratory, Department of Plant and Environmental Sciences, University of Copenhagen, 40 Thorvaldsensvej, DK-1871 Frederiksberg C,
Copenhagen, Denmark
4
Carlsberg Laboratory, 10 Gamle Carlsberg Vej, DK-1799 Copenhagen V, Denmark
*Correspondence: Birger Lindberg Møller (blm@plen.ku.dk)
http://dx.doi.org/10.1016/j.molp.2014.11.008

ABSTRACT
In recent years, biotechnology-derived production of flavors and fragrances has expanded rapidly.
The world’s most popular flavor, vanillin, is no exception. This review outlines the current state of
biotechnology-based vanillin synthesis with the use of ferulic acid, eugenol, and glucose as substrates
and bacteria, fungi, and yeasts as microbial production hosts. The de novo biosynthetic pathway of vanillin
in the vanilla orchid and the possible applied uses of this new knowledge in the biotechnology-derived and
pod-based vanillin industries are also highlighted.
Key words: metabolomics, natural products, phenylpropanoids and phenolics, molecular biology, plant biochem-
istry, synthetic biology
Gallage N.J. and Møller B.L. (2015). Vanillin–Bioconversion and Bioengineering of the Most Popular Plant Flavor
and Its De Novo Biosynthesis in the Vanilla Orchid. Mol. Plant. 8, 40–57.

INTRODUCTION possesses a pure and delicate spicy flavor, which is difficult to


duplicate by technological means.
In recent years, biotechnology-derived production of flavors and
fragrances has expanded rapidly. Vanillin, the worlds most pop- Vanillin is the most characteristic flavor compound of vanilla.
ular flavor, is no exception. This review outlines the current state Vanillin is a white crystalline powder with a pleasant, sweet, and
of biotechnology-based vanillin synthesis with the use of ferulic intense aroma, offering a vanilla-like flavor. Chemically, it is an
acid, eugenol, and glucose as substrates and bacteria, fungi, aromatic aldehyde belonging to the group of simple C6–C1
and yeasts as microbial production hosts. The de novo biosyn- phenolic compounds. Structurally, it is a phenol substituted with
thetic pathway of vanillin in the vanilla orchid and the possible an aldehyde and methoxy group at specific positions (Figure 1).
applied uses of this new knowledge in the biotechnology- Vanillin has a relatively low solubility in water at room
derived and pod-based vanillin industries are also highlighted. temperature, but is readily soluble in hot water, alcohol, and ether.

VANILLIN AND VANILLA In 2010, the annual global sales of vanillin reached more than
15 000 000 kg, with less than 1% obtained by isolation from vanilla
Vanilla is one of the most widely used flavors in the world and
pods. The production of vanilla beans and the isolation of vanillin
is applied extensively in the food, beverage, perfumery, and
from vanilla pods is a laborious and costly process (Sinha et al.,
pharmaceutical industries. Natural vanilla is a complex mixture
2008). Production of 1 kg of vanillin requires approximately
of flavors extracted from the cured pods of two different species
500 kg of vanilla pods, corresponding to the pollination of
of vanilla orchids: Vanilla planifolia and Vanilla tahitensis (Rao and
approximately 40 000 vanilla orchid flowers. The market cost of
Ravishankar, 2000). The flavor and fragrance profile of the vanilla
natural vanillin derived from vanilla pods is therefore high and
extract contains more than 200 components. Vanillin (4-hydroxy-
3-methoxybenzaldehyde) (Figure 1) is the key component, with a
concentration of 1%–2% w/w in cured vanilla pods (Sinha et al., Published by the Molecular Plant Shanghai Editorial Office in association with
2008). The vanilla extract obtained from fermented vanilla pods Cell Press, an imprint of Elsevier Inc., on behalf of CSPB and IPPE, SIBS, CAS.

40 Molecular Plant 8, 40–57, January 2015 ª The Author 2015.


Vanillin–Bioconversion and Bioengineering Molecular Plant
States (Hocking, 1997). Nevertheless, a substantial amount of
synthetic vanillin is still derived from lignin (borregaard.com).

There has been a huge surge in the exploration of more environ-


mentally friendly biosynthetic procedures to make natural flavors.
In recent years, this quest has been spurred by a growing demand
from consumers for natural products. Many consumers equate
food quality and food safety with the natural attribute. Accordingly
and guided by novel technologies, research in bioengineering of mi-
croorganisms for flavor production is rapidly expanding (Krings and
Berger, 1998; Berger, 2007). Most biotechnological approaches
for the synthesis of vanillin are based on bioconversion of certain
natural substances such as lignin, ferulic acid, eugenol, and
isoeugenol, etc., using microorganisms such as yeasts, fungi,
and bacteria as production hosts (Lesage-Meessen et al., 1996;
Hansen et al., 2009; Di Gioia et al., 2011). Plant cells and plant
tissue cultures have also been investigated for the production of
natural vanillin but are not advanced enough to be industrially
applicable (Fu et al., 1999). The continued attempts to produce
vanillin from V. planifolia tissue cultures as well as from Capsicum
frutescens cell cultures have been hampered by slow cell growth,
low gene expression levels, and poor yields (Funk and Brodelius,
1990a, 1990b; Knorr et al., 1990; Johnson et al., 1996; Rao and
Ravishankar, 2000; Dignum et al., 2001).

Microbial transformations of natural precursors to generate vanillin


Figure 1. Chemical Structure of Vanillin (4-Hydroxy-3- are accepted as ‘‘natural’’ according to the current European
Methoxybenzaldehyde). (European Directive 88/388/CEE, JO No. L184, 22 June 1988)
and US food legislation (Krings and Berger, 1998). Although US
fluctuates because of the unpredictable availability of vanilla pods. and EU regulatory requirements are somewhat similar, they do
Crop yield is tightly associated with weather conditions and the have distinct differences. Thus, the definition of a ‘‘natural’’ flavor
incidence of disease as well as local and international political according to the EU regulation is more stringent than the
and economic issues. Vanillin extracted from vanilla pods has a corresponding US regulation with respect to the source material
market price varying from around US$1200/kg to more than and the manufacturing process used. The regulatory requirements
US$4000/kg (Walton et al., 2003). The global supply of vanilla for the US and the EU ‘‘natural’’ status are quoted below:
pods is stagnant at about 2000 000 kg. Thus, the increasing
global demand for natural vanilla flavor can no longer be met The US requirements to qualify for a status as a ‘‘natural’’ flavor
with pods of the vanilla orchid as the sole source. Nowadays, are as follows: ‘‘A product derived from a source as a spice, fruit,
less than 1% of the global production of vanillin is derived from extract, oleoresin, or from a group of materials that the FDA
vanilla pods; the majority is produced synthetically using, e.g. recognizes as ‘‘natural’’ starting materials or any product process
lignin and eugenol as starting materials (Walton et al., 2000). via distillation, extraction, roasting, heating, enzymolysis, hydro-
lysis and fermentation’’ (Sabisch and Smith, 2014).
The Different Routes to Vanillin
Vanillin was isolated as the main flavor constituent of vanilla pod The EU requirements to qualify for a status as a ‘‘natural’’ flavor
extracts in 1858 by Theodore Nicolas Gobley and he subse- are as follows: ‘‘Source material must be vegetable, animal, or
quently elucidated its chemical structure. Less than 20 years after microbiological and a natural substance has to be identical to
its initial isolation, synthetically produced vanillin was marketed that found in nature and must be produced by a traditional food
(Hocking, 1997). Synthetic routes to vanillin were originally preparation process.’’ (Sabisch and Smith, 2014).
based on eugenol. Nowadays, guaiacol and lignin are the
favored starting materials (Clark, 1990). Although synthetic A number of different biotechnology-derived vanillin products have
vanillin is able to meet the global market and is rather cheap, been marketed for more than a decade (Figure 2). Rhovanil,
with a market price below 15 $/kg, chemical synthesis of produced by Solvay (previously known as Rhodia), was the first
vanillin, like many other chemical processes, has serious commercially available fermentation-derived vanillin product and
drawbacks, e.g. the use of organic solvents and hazardous was obtained by bioconversion of ferulic acid (rhodia.com). Ex-
chemicals such as sodium hydroxide in vanillin purification. turmeric vanillin is marketed by De Monchy Aromatics and is pro-
Chemical synthesis of vanillin via lignin has been calculated to duced from curcumin (demonchyaromatics.com). Sense Capture
be accompanied by the demand for safe removal of 160 kg of Vanillin is obtained by bioconversion of eugenol and marketed
waste per 1 kg of vanillin obtained. As a consequence, by Mane (Sense Capture Vanillin by Mane, 2014). De novo
concerns related to the negative environmental impact have synthesized biovanillin using glucose as a precursor will be
increased since the late 1970s and have resulted in closure of commercialized in 2014 by Evolva and International Flavors and
all lignin-derived vanillin production in Canada and the United Fragrances (Vanilla, 2014).
Molecular Plant 8, 40–57, January 2015 ª The Author 2015. 41
Molecular Plant Vanillin–Bioconversion and Bioengineering

Figure 2. Different Routes to Natural Vanillin that Are Available or Soon to Be Available on the Market.

Microbial Production of Vanillin–Bioconversion for biotechnological production of vanillin. The increasing knowl-
and Bioengineering edge of enzymes that are involved in bioconversion of ferulic
The world of microorganisms is vast and many different microbial acid and other substrates to vanillin as well as the identification
organisms are being used for efficient production of natural food and characterization of genes that are encoding them offer new
components (Krings and Berger, 1998; Berger, 2007). Promising opportunities for more targeted bioengineering of microorgan-
strains have been selected based on their growth properties, isms for vanillin production.
production potential and tolerance to high concentrations of
both product and substrate. Microorganisms and fermentation In the following sections, the substrates most commonly used
ingredients, which have been given GRAS status, are preferred. for vanillin synthesis employing microorganisms (bacteria, fungi,
GRAS is an acronym for Generally Recognized As Safe under and yeast) are presented and discussed, with emphasis on the
the regulations of the US Food and Drug Administration major issues encountered and the solutions.
(Berger, 2007).
PRODUCTION OF VANILLIN FROM
Microorganisms that are able to metabolize a range of different
precursors into vanillin have been studied experimentally. Several
FERULIC ACID
major yet common issues have challenged the successful use of Ferulic acid is the best-explored substrate for production of
microorganisms for efficient bioconversion of putative substrates vanillin. Ferulic acid is abundant in nature and shares high
into vanillin. Bottlenecks include (1) cytotoxicity of the flavor structural similarity to vanillin. Free and bound ferulic acid
products obtained and of their precursors; (2) inefficient meta- (4-hydroxy-3-methoxycinnamic acid) is one of the most abundant
bolic flow; (3) formation of undesired by-products; (4) costly phenylpropanoids in plants. In plants, ferulic acid is bio-
downstream processing methods due to the physicochemical synthesized from the aromatic amino acids phenylalanine or
properties of the substrate and the product; and (5) further meta- tyrosine (Gross and Zenk, 1969; Dewick, 1989). The two
bolism of the desired product by the selected microorganisms. precursor amino acids are produced via the shikimic acid
Bioengineering tools have been employed to circumvent pathway (Dewick, 1989). The exact route for ferulic acid
these drawbacks. Bioengineering includes the use of tools such biosynthesis in plants has not been established. When
as mutagenesis-based selection and genetic engineering for produced from phenylalanine, the first intermediate is cinnamic
enzyme/strain optimization and for cost-efficient downstream acid and the reaction is catalyzed by phenylalanine ammonia
processing. Microorganisms that exhibit rapid growth rates and lyase (Fritz et al., 1976). Subsequently, cinnamic acid 4-
are amenable to molecular genetics provide suitable platforms hydroxylase (Russell and Conn, 1967; Gabriac et al., 1991;
42 Molecular Plant 8, 40–57, January 2015 ª The Author 2015.
Vanillin–Bioconversion and Bioengineering Molecular Plant

Figure 3. Possible Bioconversion Routes of Ferulic Acid to Vanillin


(A) CoA-independent conversion of ferulic acid to vanillin via a retro-aldol reaction.
(B) CoA-dependent conversion of ferulic acid to vanillin via a retro-aldol reaction.
(C) CoA-dependent b-oxidation of feruloyl-CoA into vanillic acid.
(D) Non-oxidative decarboxylation of ferulic acid to 4-vinylguaiacol. No evidence for conversion of 4-vinylguaiacol into vanillin has been obtained.
(E) Conversion of ferulic acid into vanillin involving a reducing step.

Schoch et al., 2001) catalyzes the hydroxylation of cinnamic acid glucose, xylose, or galactose to be integrated as part of the
at the 4-position, resulting in the formation of p-coumaric acid. p- pectin or hemicellulosic fraction of cell walls. These polymers
Coumaric acid 3-hydroxylase (C3H) (Schoch et al., 2001) can be up to 50 000 Da in molecular mass (Harris and
catalyzes the hydroxylation of p-coumaric acid at the 3 Trethewey, 2010).
position, resulting in caffeic acid formation. It is not clear
whether coenzyme A (CoA) derivatives are involved or whether In complete contrast to plants, no microorganisms in nature are
the C3-hydroxylation step proceeds, e.g. through quinate and known to be able to de novo synthesize ferulic acid. Neverthe-
shikimate esters (Schoch et al., 2001). Caffeic acid could in less, a vast number of microorganisms are able to utilize ferulic
principle be O-methylated by an O-methyltransferase (Lam acid as their sole carbon energy source. In addition, when micro-
et al., 2007) to afford ferulic acid. organisms are grown on eugenol as carbon source, ferulic acid is
formed as a transient intermediate.
Ferulic acid is an important biological and structural component
of the plant cell wall and can be found free, as homodimers, or
esterified with proteins or polysaccharides in the cell wall Bioconversion of Ferulic Acid to Vanillin
(Harris and Trethewey, 2010). Ferulic acid is the precursor of As illustrated in Figure 3A–3E, ferulic acid degradation pathways
coniferyl alcohol, which provides one of the monomers for lignin in microorganisms proceed with vanillin as an intermediate. It is
biosynthesis. In cereals, ferulic acid is esterified with arabinose, this inherent capability to produce vanillin that is exploited
Molecular Plant 8, 40–57, January 2015 ª The Author 2015. 43
Molecular Plant Vanillin–Bioconversion and Bioengineering
Figure 4. The CoA-Dependent Catabolism
of Ferulic Acid Via Vanillin in Amycolatopsis
sp. 39116.

CoA-independent non–b-oxidation of ferulic


acid in microorganisms is envisioned to
occur by initial hydration of the trans-double
bond of ferulic acid, resulting in the forma-
tion of 4-hydroxy-3-methoxyphenyl-b-hy-
droxypropionic acid as a transient inter-
mediate, which by a retro-aldol reaction
is directly converted into vanillin and
stoichiometric amounts of acetic acid
(Figure 3A). The vanillin biosynthesis
pathway in the pods of V. planifolia follows
a similar route; vanillin synthase, VpVAN,
catalyzes the direct conversion of ferulic
acid and ferulic acid glucoside to vanillin
and vanillin glucoside respectively following
a retro-aldol reaction (Gallage et al., 2014).

Ferulic acid catabolism in Delftia acido-


vorans (Plaggenborg et al., 2001),
P. fluorescens AN103 (Gasson et al., 1998;
Bennett et al., 2008), Pseudomonas sp.
experimentally to obtain efficient bioconversion of ferulic acid into strain HR199 (Overhage et al., 1999), and Amycolatopsis sp.
vanillin in microorganisms. strain HR167 (Achterholt et al., 2000) is known to occur via
a CoA-dependent retro-aldol mechanism (Figure 3B). This
Five major pathways can be distinguished in microorganisms conversion is dependent on the initial activation of ferulic acid
based on the different initial reactions involved in ferulic to the corresponding CoA thioester, feruloyl-CoA. This enzymatic
acid bioconversion: (1) CoA-independent retro-aldol reaction, activation process requires CoASH, adenosine triphosphate
(2) CoA-dependent retro-aldol reaction, (3) CoA-dependent (ATP), and MgCl2 as cofactors. Formation of the ferulate-CoA is
b-oxidation, (4) non-oxidative decarboxylation, and (5) a reduc- catalyzed by 4-hydroxycinnamate-CoA (4CL) ligase, also termed
tive pathway (Figure 3A–3E). Some microorganisms have ferulate-CoA ligase and feruloyl-CoA synthetase (FCS), encoded
developed multiple pathways for bioconversion of ferulic acid. by the gene Fcs (Narbad and Gasson, 1998; Achterholt
For example, Pseudomonas fluorescens has been reported et al., 2000; Masai et al., 2002; Plaggenborg et al., 2003; Yang
to metabolize ferulic acid by decarboxylation (Huang et al., et al., 2013).The hydration of feruloyl-CoA is catalyzed by 4-
1994), reduction (Martinez-Cuesta et al., 2005), and via a hydroxycinnamoyl-CoA hydratase/lyase (HCHL) (EC 4.2.1.101)
CoA-dependent retro-aldol reaction mechanism (Gasson et al., (Mitra et al., 1999) or enoyl-CoA hydratase/aldolase (ECH),
1998). encoded by the gene Ech (Huang et al., 1994; Gasson et al.,
1998; Plaggenborg et al., 2001) (Figure 4). The 4-hydroxy-3-
The retro-aldol reaction is the reverse reaction of an aldol reaction methoxyphenyl-b-hydroxypropionyl-CoA is formed as a transient
and is well described for fatty acid degradation (Berg et al., 2012). intermediate and subsequently converted into vanillin and acetyl-
The reverse aldol reaction mechanism involves elimination of CoA by a retro-aldol mechanism.
an acetate moiety from the unsaturated ferulic acid side chain,
resulting in vanillin formation. This reaction may proceed as a Ferulic acid catabolism in Pseudomonas putida (Zenk et al., 1980)
CoA-dependent (Rosazza et al., 1995; Mitra et al., 1999) as well and Rhodotorularubra (Huang et al., 1993) is envisioned to
as a CoA-independent process (Toms and Wood, 1970; Huang proceed via CoA-dependent b-oxidation in a similar fashion to
et al., 1993; Jurkova and Wurst, 1993). fatty acid catabolism. In this catabolic scheme, ferulic acid
is initially activated to its CoA thioester, feruloyl-CoA, and
Toms and Wood (1970) reported the conversion of trans-ferulic converted to the intermediate 40 -hydroxy-3-methoxyphenyl-
acid to vanillic acid by Pseudomonas acidovorans as early as b-hydroxypropionyl-CoA. Subsequent oxidation results in the for-
1970 via a CoA-independent non–b-oxidative reaction sequence. mation of 3(40 -hydroxy-3-methoxyphenyl)-3-ketopropionyl-CoA,
In this study, administration of [2-14C]-labeled trans-ferulic which, in a CoA-consuming reaction, is cleaved to form vanillyl-
acid to resting cells of P. acidovorans resulted in the formation CoA and acetyl-CoA. Vanillyl-CoA would then be hydrolyzed to
of [2-14C]acetate and vanillic acid. A similar route for the form vanillic acid, with concomitant release of CoA-SH (Figure 3C).
degradation of ferulic acid has recently been suggested to take
place in Bacillus subtilis (Gurujeyalakshmi and Mahadevan, Non-oxidative decarboxylation of ferulic acid to 4-vinylguaiacol
1987) and in Streptomyces setonii (Sutherland et al., 1983). is a competing process known to proceed rapidly and efficiently
44 Molecular Plant 8, 40–57, January 2015 ª The Author 2015.
Vanillin–Bioconversion and Bioengineering Molecular Plant
in both bacteria, fungi, and yeast, e.g. in Bacillus coagulans offer new opportunities for the bioengineering of industrial appli-
BK07 (Karmakar et al., 2000), Delftia acidovorans (Plaggenborg cable microorganisms for vanillin biosynthesis.
et al., 2001), Enterobacter sp. Px6-4 (Li et al., 2008), Fusarium
solani (Nazareth and Mavinkurve, 1986), and Paecilomyces Recombinant strains of bacteria, fungi, and yeasts represent an
variotii (Rahouti et al., 1989) (Figure 3D). This mechanism interesting alternative to wild-type strains, and a diverse set of
occurs via the formation of a transient quinoid intermediate. methodologies have been developed to engineer and optimize
Formation of this type of quinoid intermediate from ferulic acid the bioconversion of ferulic acid to vanillin.
is enzyme promoted (Bennett et al., 2008). Studies have begun
to explore the genetic basis for non-oxidative decarboxylation
of ferulic acid. In bacteria and yeast, genes encoding phenyla- Availability of Ferulic Acid
crylic acid decarboxylase (PAD1) (EC 4.1.1.-) and ferulic acid Agricultural plant waste constitutes cheap starting materials
decarboxylase (FDC1) (EC 4.1.1.-) were identified to be essential for bioproduction. However, recovery of ferulic acid from plant
for ferulic acid decarboxylation (Cao et al., 2010; Mukai et al., cell walls appears to be somewhat complicated and costly.
2010). Impaired function of either PAD1 or FDC1 resulted in Nevertheless, this process has received considerable attention.
yeast unable to decarboxylate cinnamic acid–derived Most ferulic acid in plants is bound in the plant cell wall and re-
compounds (Mukai et al., 2010; Gallage et al., 2014). The use quires either chemical or enzymatic hydrolysis to be released.
of isotope-labeled 4-vinylguaiacol demonstrated that 4-
vinylguaiacol is subsequently converted into 4-(1-hydroxy)ethyl- Several studies have attempted to remove ferulic acid from plant
guaiacol in bacteria. Biochemical evidence for further meta- cell wall materials enzymatically (Lesage-Meessen et al., 1999;
bolism of 4-vinylguaiacol or 4-(1-hydroxy)ethylguaiacol into Bonnin et al., 2000; Zheng et al., 2007). Feruloyl esterases (EC
vanillin or other products has not been obtained (Li et al., 2008; 3.1.1.73) are enzymes that cleave the ester bonds by which
Gallage et al., 2014). ferulic acid is attached to the cell wall polymers and can be
isolated from a wide range of fungi, yeast, and bacteria (deVries
Chain shortening of the side chain of ferulic acid may also pro- et al., 1997). Fungal feruloyl esterases have been proved useful
ceed by a reductive mechanism. The proposed reaction mecha- for the hydrolysis of ferulic acid from plant materials in several
nism involves initial isomerization of ferulic acid to a transient in vitro and in vivo studies. Two feruloyl esterases, FaeA
quinoid intermediate, as described above. The quinoid intermedi- and FaeB, isolated from Aspergillus niger are able to release
ate is converted into 4-hydroxy-3-methoxy-phenylpropionic acid ferulic acid from industrial by-products such as wheat
in a reductive step and subsequently to vanillic acid (Rosazza straw, coffee pulp, apple marc, maize bran, maize fiber, etc.
et al., 1995). The exact biochemical processing of 4-hydroxy-3- (Lesage-Meessen et al., 2002; Benoit et al., 2006). Bacterial
methoxy-phenylpropionic acid to 2(40 -hydroxy-30 -methoxy- feruloyl esterases have been used for the same purpose.
phenyl)acetic acid and subsequently to vanillic acid has not Microcapsules containing bacterial feruloyl esterase expressed
been elucidated. in Lactobacillus fermentum (ATCC 11976) have been shown to
release ester-bound ferulic acid from ethyl ferulate in vitro
The conversion of ferulic acid to vanillin is, as previously (Bhathena et al., 2007).
mentioned, an intermediate step in the catabolism of ferulic
acid. Many organisms are able to rapidly catabolize the Enzymatic hydrolysis of cell walls using a combination of
vanillin formed into other products. Vanillin toxicity may have commercial polysaccharide-degrading enzymes and feruloyl
spurred evolutionary selection for rapid further conversion. In esterase has also been investigated (Faulds et al., 1997).
Pseudomonas and in the actinomycete Amycolatopsis sp. strain Currently, these methods are not economically feasible,
ATCC 39116, vanillin dehydrogenase (VDH) (EC 1.2.1.67) is able because the use of commercially available polysaccharide-
to convert vanillin to vanillic acid in a nicotinamide adenine degrading enzymes is costly and would result in significantly
dinucleotide (NAD)-dependent reaction. Vanillic acid is then increased production costs of vanillin.
catabolized to protocatechuic acid by demethylation, in which
one oxygen atom is introduced at the methyl carbon resulting Ferulic acid can also be released from plant cell walls by alkaline
in an unstable hemiacetal, which consequently decomposes treatment at high temperatures (85 C–100 C). As expected
to protocatechuic acid and formaldehyde, respectively (Priefert from the enzymatic treatments, this treatment causes release
et al., 1997; Fleige et al., 2013). This demethylation step of ester-bound ferulic acid from the cell walls (Higuchi et al.,
is catalyzed by vanillate O-demethylase, VANA and VANB (EC 1967). A recent study of alkaline hydrolysis of corncobs reports
1.14.13.82), which are encoded by the genes VanA and VanB a high yield of hydroxycinnamic acids, mainly p-coumaric
(Priefert et al., 1997; Plaggenborg et al., 2003). These two acid and ferulic acid, for potential use in vanillin production
genes are supposed to be organized in the same operon in (Torres et al., 2009). This kind of chemical release of ferulic
both Pseudomonas sp. strain HR199 and P. putida (Brunel and acid would not be considered natural processing according to
Davison, 1988; Venturi et al., 1998). the EU regulations, but would comply with registration as
‘‘natural’’ according to US legislation (Sabisch and Smith,
2014; Torres et al., 2009).
Bioengineering of Ferulic Acid–Derived Vanillin
in Microorganisms Today, ferulic acid used for commercial production of natural
Increasing knowledge of the metabolic pathways for vanillin vanillin is mostly obtained as a by-product of the production of
production via ferulic acid as well as identification and character- rice bran oil. Ferulic acid is liberated from the rice bran by
ization of the enzymes and genes involved in the individual steps enzymatic treatment to comply with the regulations for being
Molecular Plant 8, 40–57, January 2015 ª The Author 2015. 45
Molecular Plant Vanillin–Bioconversion and Bioengineering

Substrate Microorganism Length of incubation (h) Yield (g/l) References


Ferulic acid Streptomyces setonii 23 >10 Achterholt et al., 2000
Pseudomonas putida >10 Plaggenborg et al., 2003
Streptomyces sp. V-1 55 19.2 Hua et al., 2007a
Recombinant E. coli 24 5.14 Lee et al., 2009
E. coli JM 109 6 2.52 Barghini et al., 2007
Aspergillus niger and Pycnoporus 72 2.8 Zheng et al., 2007
cinnabarinus
Eugenol Pseudomonas sp. HR199 2–200 2.6 Overhage et al., 2000
Amycolatopsis sp. HR167 32 >10 Overhage et al., 2006
Isoeugenol Bacillus fusiformis SW-B9 72 32.5 Zhao et al., 2005
Recombinant E. coli BL21 (DE3) 6 28.3 Yamada et al., 2007
P. putida 24 16.1 Yamada et al., 2007
Bacillus pumilus S-1 3.8 Hua et al., 2007b
Bacillus subtilis HS8 24 1.36 Zhang et al., 2006
Psychrobacter sp. strain CSW4 1.28 Ashengroph et al., 2012
P. chlororaphis CDAE5 24 1.2 Kasana et al., 2007
Glucose Recombinant S. cerevisiae 24–168 >0.5 Hansen et al., 2013 (WO/2013/
022881); Brochado et al., 2010

Table 1. Industrially Applicable Microorganisms that Were Bioengineered to Produce Vanillin.

classified as a natural product. The cost of naturally extracted recently been re-classified as Amycolatopsis sp. ATCC 39116
ferulic acid is relatively high, with a price around US$180/kg and in a later study has been shown to yield even more
(novorate.com). With such high costs for the starting material, vanillin, with product titers as high as 13.9 g/l and a molar
production of vanillin using this approach is very costly. Key yield of 75% (Muheim and Lerch, 1999; Fleige et al., 2013).
issues related to the use of ferulic acid from agricultural waste Another Amycolatopsis species, sp. HR167, has been shown
materials as a starting material for production of natural vanillin to produce vanillin at a concentration of 11.5 g/l, with a
are thus related to the development of optimized and less molar yield of 77.8% (Rabenhorst, 1996). Such strains are
expensive production costs. thereby suitable for vanillin production on an industrial scale,
as they can tolerate both high vanillin and ferulic acid
Toxicity concentrations and display a high conversion ratio of the
Aldehydes rarely accumulate in high concentrations in biological expensive ferulic acid substrate into vanillin. However, the
systems because of their high chemical reactivity, e.g. by form- filamentous growth of actinomycetes results in highly viscous
ing Schiff bases and thereby possibly inhibiting enzymatic activ- broths, unfavorable pellet formation, and a lot of fragmentation
ity, and show toxic effects in biological systems. A major issue and lysis of the mycelium, which complicate downstream
for efficient biotechnology-derived production of vanillin is thus processing.
product toxicity. High concentrations of ferulic acid are also
toxic to many microorganisms. Toxicity is manifested by inhibi- Genes involved in ferulic acid catabolism have been heterolo-
tion of the growth of the production strain or even cell lysis gously expressed in Escherichia coli mutants with high vanillin
(Priefert et al., 2001). Microorganisms that have been shown tolerance to bypass the problems related to product toxicity.
to produce vanillin at levels exceeding 1 g/l are summarized in This includes the Fcs and Ech genes responsible for ferulic
Table 1. acid conversion in Amycolatopsis sp. HR104 (Yoon et al.,
2005). The vanillin-resistant mutant strain was obtained following
Several studies have been targeted toward isolation of bacterial NTG (N-methyl-N-nitro-N-nitrosoguanidine) mutagenesis. When
species that tolerate high concentrations of ferulic acid and grown for 48 h in medium containing 2 g/l of ferulic acid as
vanillin. One study was carried out to isolate strains from nature much as 1 g/l of vanillin was obtained. To further circumvent
that were resistant to high levels of ferulic acid and vanillin the inhibitory effect of vanillin, XAD-2 resin was used to bind
and to investigate their ability to catabolize eugenol and the vanillin formed in the medium. This increased the vanillin yield
ferulic acid (Muheim and Lerch, 1999). Actinomycetes, such as to 5 g/l in 48 h when a five-fold increase in ferulic acid substrate
Amycolatopsis sp. and S. setonii, were able to accumulate was used during incubation.
high concentrations of vanillin while at the same time exhibiting
a high tolerance toward ferulic acid. Metabolic Flow and Side Product Formation
A balanced metabolic flow in the course of conversion of sub-
S. setonii was able to produce vanillin, with levels reaching strates into vanillin and prevention of formation of unwanted
6.4 g/l in shake flask experiments. The same strain has side products are other important factors in the production of
46 Molecular Plant 8, 40–57, January 2015 ª The Author 2015.
Vanillin–Bioconversion and Bioengineering Molecular Plant
vanillin. The flow of metabolites in the conversion of ferulic acid Another approach to reduce by-product formation is to use a
into vanillin varies among different bacterial strains. Pseudo- two-step fermentation process involving two different microbial
monas strains have been found to be excellent converters of organisms. In one such case, ferulic acid catabolism in A. niger
ferulic acid to vanillin (Muheim and Lerch, 1999). However, the and vanillin metabolism in Pycnoporus cinnabarinus were com-
vanillin formed is readily oxidized into vanillic acid and/or bined. In the first step, ferulic acid was catabolized to vanillic
reduced to vanillyl alcohol. In contrast, S. setonii metabolizes acid in high yield by the micromycete A. niger. In the second
ferulic acid to vanillin as an overflow product because of a step, the vanillic acid formed was reduced to vanillin by the
bottleneck in the oxidation of vanillin to vanillic acid catalyzed basidiomycete P. cinnabarinus. The vanillic acid and vanillin titers
by vanillin dehydrogenase (VDH). When high concentrations of obtained were 920 mg/l and 237 mg/l, respectively (Lesage-
ferulic acid are provided, the activity of ferulic acid–degrading Meessen et al., 1996).
enzymes is much higher in this strain compared with VDH,
directing the metabolic flow toward accumulation of vanillin. The influence of the use of different genetic approaches to achieve
This makes S. setonii a good candidate for the production of high gene expression including plasmid copy number and pro-
vanillin on an industrial scale. moters regulating genes involved in vanillin synthesis has been
studied in recombinant E. coli. When the E. coli strain JM109
Several approaches have been taken to impair the undesired con- was engineered with a low copy number vector pBB1 carrying
version of vanillin into vanillic acid. In Pseudomonas strains, this the Fcs and Ech genes isolated from P. fluorescens BF13
was explored by mutating the vdh gene encoding VDH (Di Gioia (Barghini et al., 2007), a final vanillin concentration of 2.52 g/l
et al., 2011). A higher concentration of vanillin accumulated in was obtained after 6 h of incubation by sequential induction with
recombinant P. putida BO14, which was engineered to have a 1.1 mM ferulic acid at resting cell conditions. To further improve
defective vdh gene (Okeke and Venturi, 1999). In the case of vanillin formation, an integrative vector pFR2 was constructed
P. fluorescens BF13, in which expression of the vdh gene was carrying the Fcs and Ech genes stably integrated into the lacZ
blocked, a complete and simultaneous loss of the ability to gene of E. coli. The recombinant strain was stable and
metabolize ferulic acid and vanillin was observed. The effect more efficient in vanillin synthesis compared with the strain
was reversed on introduction of multiple copies of the Fsc gene. expressing genes encoding ferulic acid catabolizing enzymes
This gene encodes feruloyl aldehyde dehydrogenase, which from a low copy vector. The recombinant strain produced 6.6 kg
catalyzes the activation of ferulic acid to the corresponding CoA vanillin per 1 kg biomass in resting cells (Ruzzi et al., 1997).
ester, feruloyl-CoA. The recombinant P. fluorescens BF13 was
able to utilize ferulic acid and yielded up to 1.28 g/l vanillin in a PRODUCTION OF VANILLIN FROM
stirred tank reactor after 8 h (Di Gioia et al., 2011). A mutant of
EUGENOL AND ISOEUGENOL
P. fluorescens 103, with impaired VDH activity (vdh), was unable
to utilize ferulic acid in a similar manner to the wild-type A major drawback of ferulic acid–based vanillin production in
P. fluorescens BF1. It was suggested that this effect was due microorganisms is the high cost of ferulic acid. Eugenol and
to the inactivation of 4-hydroxycinnamate-CoA ligase, 4CL isoeugenol may be used as alternative substrates for vanillin for-
(Martinez-Cuesta et al., 2005), which is also known to catalyze mation because oxidative metabolism of the two compounds by
the activation of ferulic acid to the corresponding CoA ester, microorganisms in nature proceeds with ferulic acid and vanillin
feruloyl-CoA. These observations indicate that inactivation of as intermediates. Eugenol (2-methoxy-4-(2-propenyl)-phenol) is
the vdh gene affects the expression of other upstream genes the principal component of clove oil prepared from the clove
that are essential for catabolism of ferulic acid in Pseudomonas tree, Syzgium aromaticum, and is isolated on an industrial scale
strains; especially the reaction of ferulic acid activation to (Bauer et al., 2008). The market price is as low as US$5/kg for
feruloyl-CoA seems to be compromised. clove oil and around US$50/kg for isolated eugenol. The
starting materials for vanillin production based on eugenol are
The mutation-based approaches undertaken to impair undesired thus a lot cheaper for biotechnology-derived vanillin production
oxidation of vanillin into vanillic acid are challenging because than ferulic acid. Eugenol is considered a GRAS compound
some microorganisms possess more than one Vdh gene. vdh when used as a food additive and accordingly has developed
mutants of Pseudomonas sp. strain HR199 and P. putida into a popular precursor for industrial production of vanillin using
KT2440 were able to oxidize vanillin, indicating the existence of microorganisms (fda.gov).
additional VDH-like enzymes with overlapping functional activ-
ities. Accordingly, microorganisms that lack the ability to convert Bioconversion of Eugenol to Ferulic Acid and Vanillin
vanillin into vanillic acid when the Vdh gene is functionally The oxidative catabolism of eugenol has been studied in a num-
impaired are good candidates for bioengineering. For example, ber of different microorganisms. In Pseudomonas, the enzymes
the vdh-deletion mutant of Amycolatopsis sp. ATCC 39116 re- and the corresponding structural genes have been identified.
sulted in a two to three times higher yield of vanillin than the The catabolism of eugenol in Pseudomonas strains proceeds
wild-type strain. It was demonstrated that the Amycolatopsis sequentially via coniferyl alcohol, coniferyl aldehyde, ferulic
sp. ATCC 39116, VDH, specifically catalyzes vanillin oxidation acid, vanillin, and vanillic acid to protocatechuic acid. Protocate-
without being involved in ferulic acid activation (Fleige et al., chuic acid is further catabolized by ortho cleavage (Tadasa, 1977;
2013). The oxidation of vanillin to vanillic acid was strongly Priefert et al., 1997; Achterholt et al., 1998; Overhage et al., 1999;
reduced when Vdh gene was impaired in Amycolatopsis sp. Priefert et al., 1999).
ATCC 39116. Consequently, Amycolatopsis sp. ATCC 39116 is
a suitable microorganism to be optimized for industrial vanillin Eugenol degradation in Pseudomonas involves two oxidation
production by further bioengineering. steps. The initial step is the bioconversion of eugenol into
Molecular Plant 8, 40–57, January 2015 ª The Author 2015. 47
Molecular Plant Vanillin–Bioconversion and Bioengineering
Figure 5. Eugenol Degradation Pathway in
Pseudomonas.

formed would then be catabolized to


vanillin, vanillyl alcohol, and vanillic acid, as
described above in ferulic acid metabolism.

Despite the well-characterized eugenol


degradation pathways mentioned above,
there is evidence for the existence of other
possible eugenol degradation pathways in
microorganisms.

A potential route for oxidative conversion of


eugenol into ferulic acid is via the action of
vanillyl alcohol oxidases. Vanillyl alcohol
oxidase (EC 1.1.3.7) (VAO) is a type of flavo-
protein oxidase that was first isolated from
Penicillium simplicissimum grown on ve-
ratryl alcohol (3,4-dimethoxybenzylalcohol)
(Benen et al., 1998). Enzymes of this family
are known to be active on a range of
phenolic compounds, including eugenol
and vanillyl alcohol. The enzyme has been
shown to catalyze the oxidation of eugenol
to coniferyl alcohol and vanillyl alcohol to
vanillin (Overhage et al., 2003). Genes
encoding VAO-like enzymes are known
both from bacteria, e.g. Rhodococcus
(Jin et al., 2007) and fungi (Lambert et al.,
2014). A recombinant Amycolatopsis
sp. HR167 gained the ability to utilize
eugenol following expression of the vanillyl
coniferyl alcohol. This step is catalyzed by the heterodimeric alcohol oxidase gene (vaoA) from P. simplicissimum together
enzyme eugenol hydroxylase (EC 1.14.15.-) encoded by the with the coniferyl ADH (calA) and CALDH (calB) genes from
genes EhyA and EhyB (Figure 5). The two subunits, EhyA and Pseudomonas sp. HR199 (Overhage et al., 2006).
EhyB, constitute an enzyme of the flavocytochrome c class, in
which the gene EhyA encodes the cytochrome c subunit The capacity for eugenol bioconversion and ferulic acid catab-
(Priefert et al., 1999). The two genes were characterized in olism in Rhodococcus sp. I24 and Rhodococcus sp. PD630
Pseudomonas sp. strain (HR199) and homologs of EhyA and shows significant differences. In contrast to Rhodococcus sp.
EhyB were also identified in a few other Pseudomonas strains PD630, Rhodococcus sp. I24 can tolerate up to 2.5–3.0 mM
such as P. putida and P. fluorescens E118, which utilize eugenol, implying a natural occurrence of effective eugenol
eugenol as their sole carbon energy source (Priefert et al., catabolism in this strain. Genomic sequence analysis of Rhodo-
1999). Functional expression of the two genes in various coccus sp. I24 did not reveal significant sequence identity to
Pseudomonas strains that were unable to utilize eugenol genes encoding known eugenol hydroxylases such as the
showed that expression of EhyA and EhyB rendered the strain EhyA-B from other bacterial strains mentioned above. This
able to grow on eugenol (Priefert et al., 1999). This provided would indicate that the initial steps of degradation of eugenol
independent confirmation that EhyA and EhyB catalyze the in Rhodococcus sp. I24 are catalyzed by an enzyme or several
initial step in oxidative eugenol degradation. enzymes that differ from those in pseudomonad strains. Alterna-
tively, eugenol catabolism in Rhodococcus sp. I24 may be due
In the second oxidative step, coniferyl alcohol is oxidized to to VAO activity.
coniferyl aldehyde. This step is catalyzed by coniferyl alcohol
dehydrogenase (ADH) (EC 1.1.1.94) encoded by the gene CalA. In contrast to the well-described eugenol catabolism in bacteria,
Coniferyl aldehyde is subsequently converted to ferulic acid by eugenol degradation pathways in yeast and fungi are less inves-
coniferyl aldehyde dehydrogenase (CALDH) (EC 1.2.1.68) en- tigated. The expression of the vaoA gene from P. simplicissimum
coded by the gene CalB (Achterholt et al., 1998; Plaggenborg in wild-type Saccharomyces cerevisiae resulted in conversion of
et al., 2006) (Figure 5). Both enzymes are dependent on NAD+ administered eugenol to coniferyl alcohol. S. cerevisiae has
as cofactor and their structural genes CalA and CalB were been reported to convert coniferyl alcohol to ferulic acid in nature
identified in Pseudomonas sp. strain (HR199). The ferulic acid due to a dehydrogenase enzyme activity (Lambert et al., 2014).
48 Molecular Plant 8, 40–57, January 2015 ª The Author 2015.
Vanillin–Bioconversion and Bioengineering Molecular Plant
A fed-batch bioconversion process from eugenol to coniferyl eugenol- and isoeugenol-derived vanillin production encounters
alcohol by the fungus Byssochlamys fulva V107 has been re- similar drawbacks as described for ferulic acid. Consequently,
ported to yield 21.9 g/l coniferyl alcohol within 36 h with a molar only two major issues and the solutions to those are highlighted
yield of 94.6% (Furukawa et al., 1999). Likewise, bioconversion here; namely, the toxicity caused by eugenol and isoeugenol
of eugenol to 4-vinylguaiacol was observed in Fusarium solani, and the issue of side product formation.
with ferulic acid implied to be formed as an intermediate in the
conversion, although ferulic acid could not be detected in the Toxicity
medium (Nazareth and Mavinkurve, 1986).
Both eugenol and isoeugenol are synthesized in plants as antimi-
crobial compounds and are toxic at high concentrations
Ferulic acid formed via eugenol catabolism in fungi and yeast
(Karapinar, 1990; Koeduka et al., 2006). For that reason,
would be a direct substrate for the synthesis of natural vanillin.
microorganisms have also developed various degradation
S. cerevisiae does not process the ability to convert ferulic
pathways to maintain the level of eugenol and isoeugenol
acid to vanillin and is known to decarboxylate ferulic acid
below a threshold concentration to avoid toxicity and enable
efficiently to 4-vinylguacol. Nevertheless, a bioengineered
proliferation. To circumvent intoxication of the production
pathway could be envisioned using recombinant S. cerevisiae
strains used for vanillin production caused by eugenol and
mutants, in which the pad1fad1 genes encoding phenylacrylate
isoeugenol, microbial strains with high tolerance to these
decarboxylase (PAD1) and ferulate decarboxylase (FDC1)
compounds were obtained by screening and bioengineering
are disrupted, thus preventing conversion of ferulic acid
tools were employed to modify the genetic makeup of these
to 4-vinylguaiacol (Benen et al., 1998). Introduction of the
strains. One such example is the use of recombinant E. coli
gene encoding vanillin synthase (VpVAN) from the orchid
BL21(DE3) cells, which do not possess vanillin-degrading activity
V. planifolia would catalyze the conversion of ferulic acid
and express the IEM of P. putida IE27 (Yamada et al., 2008).
and ferulic acid glucoside to vanillin and vanillin glucoside,
These recombinant E. coli cells were able to efficiently produce
respectively (Gallage et al., 2014).
vanillin from isoeugenol without forming vanillic acid as a by-
product.
Bioconversion of Isoeugenol to Vanillin
Like eugenol, isoeugenol is also present in the essential oil of Side Product Formation
the clove tree (Syzygium aromaticum). Bioconversion of isoeu-
As mentioned previously, further degradation of the vanillin
genol to vanillin in both bacteria and fungi has been proposed
formed into by-products constitutes a major challenge in
to occur via an epoxide-diol pathway involving oxidation of
biotechnology-based vanillin production in microorganisms. In
the side chain of isoeugenol. Isoeugenol monooxygenase
eugenol-catabolizing bacteria, two possibilities have been inves-
(IEM) encoded by the gene Iem is the enzyme responsible for
tigated to circumvent this problem. These are based on prevent-
catalyzing the conversion of isoeugenol to vanillin. The final
ing vanillin degradation by increased pathway flux toward vanillin
bioconversion product from isoeugenol is the corresponding
formation and on continuous removal of the product, vanillin, by
substituted benzoic acid.
different product removal techniques.
Several studies report industrially viable vanillin production from
Increased production of vanillin by over-expression of IEM was
isoeugenol using bacteria and fungi (Shimoni et al., 2000, 2002;
studied in isoeugenol and eugenol metabolizing Pseudomonas
Ashengroph et al., 2010) employing strains tolerant to high
mitroreducens, which is known to contain the Iem gene and its
levels of isoeugenol (Yamada et al., 2007). P. putida IE27 and
regulatory protein IemR. IemR was suggested to be a positive
Bacillus fusiformis have been reported as efficient converters of
transcriptional regulator of Iem. It was suggested that the high
isoeugenol into vanillin (Kasana et al., 2007) with B. fusiformis
level expression and use of multiple copies of Iem and IemR re-
yielding 32.5 g/l vanillin after 72 h incubation. The P. putida
sulted in increased production of vanillin (Ryu et al., 2012) by
IE27 strain produces 16.1 g/l of vanillin following 24 h
increasing the metabolic flux toward vanillin.
incubation. The high vanillin production was obtained from
continuous addition of isoeugenol to the cultures, which helps
A few attempts have been made to study the potential advantage
to prevent further oxidation of the vanillin formed into vanillic
of using various product removal techniques to prevent vanillin
acid (Shimoni et al., 2000, 2002; Zhang et al., 2006; Kasana
degradation. These involve the use of different resins for product
et al., 2007; Yamada et al., 2007; Ashengroph et al., 2010).
removal. In cultures of P. cinnabarinus grown at bioreactor scale,
Isoeugenol bioconversion to vanillin has also been observed in
improved transformation of vanillic acid into vanillin was attemp-
the fungus, A. niger ATCC 9142 (Abraham et al., 1988). In this
ted by binding the toxic vanillin to added absorbents (Topakas
study, the yield of vanillin was considerably lower due to further
et al., 2003). The utilization of cellobiose and adsorbent resins
catabolism of vanillin into vanillyl alcohol and vanillic acid.
such as Amberlite XAD-2 and Diaion HP20 resulted in only limited
Nonetheless, currently, the yields of vanillin produced from
improvements in the product recovery process (Stentelaire et al.,
isoeugenol are higher than those obtained from eugenol (see
1998; Bonnin et al., 2000). In contrast, the use of macroporous
Table 1).
adsorbent resins like DM11 gave promising results in fed-batch
biotransformation of ferulic acid using Amycolatopsis sp. ATCC
Bioengineering of Eugenol- and Isoeugenol-Derived 39116 and S. setonii, resulting in vanillin yields of 12 g/l. High
Vanillin Production in Microorganisms concentrations of vanillin, above 10 g/l at 20 C, result in precip-
Eugenol- and isoeugenol-derived vanillin production in microor- itation of crystalline vanillin and offer a convenient method of
ganisms proceeds via ferulic acid as an intermediate. Thus, isolation.
Molecular Plant 8, 40–57, January 2015 ª The Author 2015. 49
Molecular Plant Vanillin–Bioconversion and Bioengineering
Figure 6. Engineered De Novo Vanillin
Biosynthesis from Glucose in the Yeasts
S. cerevisiae and Schizosaccharomyces
pombe.

phenomenon was suggested to be due


to soil bacteria such as Pseudomonas
strains metabolizing glucose to vanillin and
vanillic acid (Ryu et al., 2012). However,
no direct evidence was provided to
demonstrate that the vanillin observed was
actually synthesized from glucose by the
Pseudomonas studied. The trace amounts
of vanillin formed most likely arose from
microbial degradation of other phenolic
compounds such as ferulic acid, eugenol,
or lignin, which would also have been
present in the soil sample.

Biotechnological Production of
Vanillin from Glucose
Despite no documented evidence for direct
bioconversion of glucose into vanillin by
microorganisms in nature, biosynthesis of
vanillin from glucose has been demon-
strated using both recombinant E. coli and
yeasts.

Li and Frost (1998) devised a route for


microbial production of vanillin from
glucose, in which de novo biosynthesis of
PRODUCTION OF VANILLIN FROM vanillic acid in E. coli was combined with enzymatic in vitro
SUGARS conversion of vanillic acid to vanillin. The recombinant E. coli
KL7 was engineered to dehydrate 3-dehydroshikimic acid to
Plants and photosynthetic microorganisms convert energy protocatechuic acid (3,4-dihydrobenzoic acid) by the action
from sunlight into chemical energy in the form of ATP and of 3-dehydroshikimic dehydratase (3DSD) (EC 4.2.1.118) en-
NADPH. The photosynthetic processes thus drive the formation coded by the gene AroZ from the dung mold fungus Podospora
of glucose and other sugars from H2O and CO2. The abundance anserina. 3-Dehydroshikimic acid is an intermediate in the shiki-
and low cost of sugars make them very attractive substrates mate pathway resulting in biosynthesis of aromatic amino acids.
for microbial production of chemicals like vanillin. Glucose is Protocatechuic acid was then converted to vanillic acid by a
the most explored sugar substrate for vanillin biosynthesis human catechol-O-methyltransferase (COMT) (EC 2.1.1.6).
because it is cheap and readily utilized by most microorganisms. Reduction of vanillic acid to vanillin was carried out in vitro using
a cellular extract of Neurospora crassa, exhibiting the required
The cost of glucose can be as low as US$0.30/kg. It is the cheap- aromatic carboxylic acid reductase (ACAR) activity (Gross and
est substrate used for vanillin production by bioengineering Zenk, 1969; Li and Frost, 1998). The use of recombinant E. coli
(Hansen et al., 2009). It is also valuable as a cheap primary KL7 for vanillin production from glucose on an industrial scale
energy source for the production strain. Moreover, glucose is a according to the outlined approach has not yet been achieved.
more attractive substrate in comparison with ferulic acid, A main obstacle is the costly in vitro step, which is dependent
eugenol, and other phenolic compounds, because it is not toxic on ATP, NADPH, and Mg2+ supplementation.
to microorganisms and does not pose any problem with
potential off-taste, which is known to be associated with the Incorporation of the entire vanillin biosynthetic pathway within
bioengineering approaches to produce ‘‘natural’’ vanillin using, a single microorganism would be much more attractive, because
e.g. eugenol or ferulic acid (Evolva, personal communication). it would circumvent the demand for supplementation with
expensive cofactors. In 2009, this was achieved by Hansen
Bioconversion of Glucose to Vanillin et al. (2009), who demonstrated the first example of microbial
Biochemical or genetic evidence for direct bioconversion of vanillin biosynthesis from glucose in the yeasts, S. cerevisiae
glucose to vanillin in any naturally occurring microorganism and Schizosaccharomyces pombe. The vanillin biosynthesis
has not been provided. In 1966, glucose-induced vanillin for- pathways engineered into these yeasts are outlined in Figure 6.
mation in soils was reported (Kunc and Macura, 1966). This The pathway has similarities to the pathway demonstrated by
50 Molecular Plant 8, 40–57, January 2015 ª The Author 2015.
Vanillin–Bioconversion and Bioengineering Molecular Plant
Li and Frost in 1998, but a major difference and breakthrough is for in vivo microbial production of vanillin from glucose resulted
that it encompasses the full in vivo conversion of glucose to in accumulation of protocatechuic acid, vanillic acid, and isova-
vanillin within a single microorganism. This was achieved by nillic acid, as the main products obtained (Li and Frost, 1998).
introduction of an ACAR (EC 1.2.1.30) from Nocardia iowensis, This implied a bottleneck at the methylation step, caused by
in combination with a phosphopantetheinyl transferase, which insufficient COMT enzyme activity as required to convert
in S. cerevisiae is required for proper activation of the ACAR protocatechuic acid into vanillic acid in E. coli KL7.
enzyme (Venkitasubramanian et al., 2007). The ACAR genes
catalyze the ATP- and NADPH-driven reduction of protocate- The de novo vanillin biosynthesis pathways constructed in yeast
chuic acid to protocatechuic aldehyde and of vanillic acid into and E. coli were based on the use of a COMT enzyme that also
vanillin based on endogenously produced cofactors in the yeasts. catalyzed an unwanted O-methylation at the para position of
protocatechuic acid, resulting in isovanillic acid formation,
The pathway demonstrated in yeast utilizes the gene encoding compared with the desired methylation at the meta position, giv-
3DSD from P. anserina to catalyze the formation of protocatechuic ing rise to vanillic acid. Isovanillic acid may in turn be reduced
acid from 3-dehydroshikimate. From protocatechuic acid, the to isovanillin but not to vanillin. Isovanillin is a flavor component
pathway may then proceed via vanillic acid formed by O-methyl- and is an unwanted side product in the biotechnology-based
ation by the human COMT (EC 2.1.1.6) (Hansen et al., 2009) and vanillin industry. Separation of isovanillin from the desired product
subsequent reduction to vanillin by ACAR. Alternatively, vanillin is expected to be rather complicated as vanillin and isova-
protocatechuic aldehyde formed by reduction of protocatechuic nillin share similar physicochemical properties. To circumvent this
acid by the ACAR enzyme may be O-methylated into vanillin by issue, Hansen et al. screened a number of different O-methyl-
the COMT. The de novo vanillin biosynthesis platform is now in transferases (OMTs) to identify an OMT that was specific for the
the process of scaling up to industrial needs and will be on methylation of the 3-hydroxy group of protocatechuic aldehyde
market under the label ‘‘natural vanillin’’ (Vanilla, 2014). and not able to methylate the 4-hydroxy group. In an alternative
approach, Hansen et al. introduced site-specific mutations which
Nonetheless, glucose-based vanillin biosynthesis pathways were selected based on 3-dehydroshikimic enzyme structure
constructed in recombinant E. coli and yeasts have several chal- prediction. By these efforts, the group managed to construct
lenges beyond just getting the product formed. Such challenges mutant versions of the human COMT carrying either an L198Y
are related to formation of unwanted side products, balanced or an E199A mutation, which made the mutated enzyme almost
efficiency of the enzymatic steps involved, and toxicity of precur- entirely specific for methylation at the meta position, thus
sors or end product. avoiding undesired isovanillin production as well as achieving a
higher turnover of precursors (Hansen et al., 2013).
Toxicity
To improve the metabolic flux toward a higher yield of vanillin in
As pointed out previously, vanillin is toxic to living cells in high con-
the de novo vanillin biosynthetic pathway in yeast, mutations in
centrations. Dealing with this issue is an important prerequisite for
the production strains were carried out. A mutation was intro-
building economically viable biotechnology-derived vanillin cell
duced in the AROM enzyme complex (ARO1) to increase the
factories. In the case of S. cerevisiae, vanillin production beyond
accumulation of 3-dehydroshikimate (Hansen et al., 2013). This
0.5–1 g/l was toxic to the yeast, as shown by hampered growth
mutation resulted in increased accumulation of protocatechuic
and biosynthesis (Hansen et al., 2009). The natural vanillin
acid and thereby redirected the metabolic flux from aromatic
biosynthesis pathway in the vanilla orchid, V. planifolia, has an
amino acid synthesis to vanillin precursor production. As a
elegant solution to cope with the toxicity issue by glucosylation
consequence, subsequent identification and isolation of a more
of vanillin to vanillin-b-d-glucoside. The same strategy was
efficient and more specific ACAR enzyme, which was able
implemented by Hansen et al. (2009), in which A. thaliana uridine
to catalyze the conversion of the high concentrations of
diphosphate–glucose glycosyltransferase (UGT), UGT72E2 was
protocatechuic acid to protocatechuic aldehyde, arose as a
employed to glycosylate vanillin to produce the less toxic
key factor. The group was able to isolate an efficient ACAR
vanillin-b-glucoside as the final product. Hansen et al. reported
gene from Neurospora crassa which, when expressed in the
that extracellular concentration of vanillin-b-d-glucoside even
yeasts efficiently catabolized the high concentrations of
above 25 g/l has no effect on yeast growth (Hansen et al., 2009).
protocatechuic acid. This was one of the key features for the
Vanillin-b-D-glucoside has higher water solubility than vanillin
successful generation of recombinant S. pombe and S.
and could very well also be considered to provide a sink that
cerevisiae capable of de novo synthesizing vanillin (Hansen
can aid in directing the pathway toward vanillin synthesis.
et al., 2013) (Evolva, personal communication).
Moreover, glycosylation of vanillin to vanillin glucoside can also
be achieved by vanillin-specific UGT, VpUGT72U1 isolated from
As described in previous sections, further degradation of vanillin
the vanilla pods of V. planifolia (Gallage et al., 2014).
to vanillic acid and vanillyl alcohol is a major drawback in microbial
vanillin production. To circumvent this issue, S. cerevisiae ADH
Metabolic Flow and Side Product Formation enzymes became a target of modification for optimized vanillin
A major issue with developing glucose-based vanillin biosyn- production in yeast. The activity of the ADH enzymes is directly
thesis pathways in bacteria and yeast is the possible inefficient related to the concentration of dissolved oxygen, the carbon
or unbalanced activities of enzymes introduced into the produc- source, and ethanol. This is because the primary function of the
tion strains. Inefficiency of specific enzymes gives rise to bottle- ADH enzymes is to oxidize ethanol into acetaldehyde when oxida-
necks, resulting in the escape and accumulation of intermediates tion of pyruvate through the tricarboxylic acid cycle is hampered
and reduced amounts of the final product. The first reported route due to lack of oxygen in the mitochondria. In total, 29 known or
Molecular Plant 8, 40–57, January 2015 ª The Author 2015. 51
Molecular Plant Vanillin–Bioconversion and Bioengineering
hypothesized ADHs, aryl-ADHs, and related aldehyde reductases orchid. This situation has now changed. Gallage et al. (2014)
were tested for their ability to reduce vanillin to vanillyl alcohol in recently demonstrated that de novo biosynthesis of vanillin in
mutants of S. cerevisiae (Hansen et al., 2009). From the palette the orchid V. planifolia is catalyzed by a single enzyme, vanillin
of tested enzymes, ADH6 was recognized as the most important synthase (VpVAN).
gene encoding vanillin reductase activity in S. cerevisiae. The
adh6 mutants in S. cerevisiae grew normally under all growth VpVAN catalyzes the two-carbon cleavage of ferulic acid and its
conditions and showed a 50% decrease in converting vanillin to glucoside into vanillin and vanillin glucoside, respectively. The
vanillyl alcohol (Hansen et al., 2009). The efforts to prevent enzyme is strictly specific to the substitution pattern at the aro-
oxidation of vanillin into vanillic acid using genetic engineering matic ring. The substrate specificity of the enzyme was demon-
have had only limited success to date. strated in vitro by coupled transcription/translation assays and
in vivo by stable expression in yeast. VpVAN was found to be
OTHER SUBSTRATES FOR MICROBIAL localized in the inner part of the vanilla pod and high transcript
levels were observed in single cells located a few cell layers
VANILLIN PRODUCTION from the inner epidermis. Transient expression of VpVAN in
Lignin tobacco and stable expression in barley in combination with
The suitability of lignin as a substrate for vanillin production on an the action of endogenous ADHs and UGTs resulted in vanillyl
industrial scale has been investigated. Lignin is a constituent of alcohol glucoside formation from endogenously present ferulic
the plant cell wall and the second most abundant plant polymer acid. Vanillyl alcohol glucoside was obtained as the final product
in the world, as well as one of the most abundant natural sources because vanillin is readily reduced to its alcohol and glycosylated
of aromatic compounds. Although lignin is a natural polymer, in plants and microorganisms. Ground ivy (Glechoma hederacea)
enzymatic degradation of lignin into its structural monomers is also produces vanillin. A gene encoding an enzyme showing
not possible. Vanillin production from lignin on an industrial scale 71% sequence identity to VpVAN was identified and shown by
is therefore based on harsh chemical oxidation. This chemically transient expression in tobacco to be a vanillin synthase.
synthesized vanillin is marketed as ‘‘synthetic vanillin’’. Several
studies have explored biological degradation of lignin using The conversion of ferulic acid and its glucoside into vanillin and
white-rot fungi. During biological degradation of lignin, vanillin the corresponding glucoside is envisioned to proceed sequen-
was formed in trace amounts. However, characterization of the tially by an initial hydration addition reaction followed by a CoA-
enzymes that are involved in lignin degradation, resulting in independent retro-aldol elimination reaction, resulting in the for-
vanillin and the optimization of such pathways for industrial mation of vanillin with concomitant release of stoichiometric
biotechnology-derived vanillin production, is incomplete (Kirk amounts of acetic acid. This reaction is similar to ferulic acid
and Farrell, 1987; Priefert et al., 2001; Martinez et al., 2004). degradation in microorganisms, as described previously.

VpVAN shows high amino acid similarity to enzymes within the


Aromatic Amino Acids and Phenolic Stilbenes
cysteine protease family (Garcia-Lorenzo et al., 2006; Gallage
Bioconversions of stilbenes and aromatic amino acids to vanillin et al., 2014). The cysteine protease family encompasses a large
have also been studied. Stilbenes are commonly found in spruce group of enzymes with versatile physiological functions and a
bark and can be oxidatively cleaved to the corresponding aro- broad range of substrate specificities. The cysteine proteases
matic aldehydes. The cleavage reaction is catalyzed by lignostil- identified in plants are also diverse. Plants may contain more
bene-a,b-dioxygenases. Genes encoding stilbene-degrading than 150 different cysteine proteinases, with the papain-like
enzymes from Pseudomonas paucimobilis strain TMY 1009 cysteine proteinases as one of the major families encompassing
have been cloned and expressed in E. coli (Priefert et al., 2001). 30–50 different members dependent on the plant species
(Garcia-Lorenzo et al., 2006). The function of the individual
Phenylalanine is the main precursor for biosynthesis of phenyl- cysteine proteinases is in general not well characterized; papain
propanoids such as flavonoids, lignin, coumarines, and stilbenes, (EC 3.4.22.2) from the latex of Carica papaya is one of the
and some of the enzymes and genes responsible for their for- exceptions (Otto and Schirmeister, 1997). In general, cysteine
mation are known. The metabolism of l-phenylalanine has proteases are expressed as pre-proteins with an N-terminal
been studied in detail in white-rot fungi (Casey and Dobb, 1992; endoplasmic reticulum (ER) - targeting signal peptide as part of
Jensen et al., 1994; Krings et al., 1996). Yet, none of these the pro-peptide domain, which comprises 130–160 amino acid
phenylpropanoids yielded considerable amounts of vanillin; residues (Cambra et al., 2012). The pro-peptide sequence is
therefore, these classes of compounds are not used for removed either by a processing enzyme or by autocatalytical
industrial-scale synthesis of vanillin at the present time. processing, thereby generating the mature cysteine proteinase
(Turk et al., 2012). Two putative protease cleavage sites in
DE NOVO BIOSYNTHESIS OF VANILLIN VpVAN were identified after residues 61 (RFAR/RYGK) and 137
IN THE VANILLA ORCHID VANILLA (VDGV/LPVT) (Gallage et al., 2014). In vitro transcription/
translation experiments showed no evidence of autocatalytic
PLANIFOLIA processing of the VpVAN protein formed and demonstrated
The option to establish biotechnological production of vanillin that removal of the pro-peptide requires the action of a separate
based on heterologous expression of the genes encoding the en- processing enzyme (Gallage et al., 2014). The pro-peptide
zymes catalyzing the synthesis of vanillin in the native vanilla sequence may control intracellular targeting, promote proper
orchid has not yet been tested for the simple reason that the folding of the mature enzyme, and/or serve to maintain the
pathway and genes involved had not been identified in the vanilla enzyme in an inactive form in the cell to balance its function
52 Molecular Plant 8, 40–57, January 2015 ª The Author 2015.
Vanillin–Bioconversion and Bioengineering Molecular Plant
according to physiological demands (Otto and Schirmeister, lactic acid bacteria and baker’s yeast (S. cerevisiae) are popular
1997). A modified VpVAN sequence in which the VpVAN pro- choices from the consumer perspective.
peptide was replaced with the putative ER-targeting putative
signal peptide and the putative pro-peptide protease cleavage An entirely new opportunity for biotechnology-based production
site from the tobacco cysteine protease that showed the highest of natural vanillin may arise from the recent identification of the
sequence identity to VpVAN was transiently expressed in to- vanillin synthase enzyme VpVAN from the vanilla orchid Vanilla
bacco. The modifications resulted in the formation of increased planifolia and from ground ivy (Glechoma hederacea) (Gallage
levels of vanillyl alcohol glucoside in comparison with parallel ex- et al., 2014). As described, vanillin and vanillin glucoside
periments with VpVAN. biosynthesis in the vanilla orchid proceeds with ferulic acid and
its glucoside as precursors and is catalyzed by a single enzyme
that does not require the presence of specific cofactors. If high
FUTURE PERSPECTIVES expression levels can be obtained in yeast production strains
Industrial application of bioengineered microorganisms for vanillin and the enzyme is stable and has proper kinetic characteristics,
production has gained a lot of attention not only from the flavor this may constitute an alternative to the current yeast
and fragrance industries but also from environmental groups, production systems. Similar to some of the other yeast-based
the general public, and politicians. The current and future impor- production systems, such a production route would be depen-
tance of biotechnology-based vanillin production is apparent dent on the supply of rather costly ferulic acid as substrate.
from the vast amount of scientific papers published on this topic
and from the number of patents that have been applied for and The identification of vanillin synthase as a hydratase/lyase-type
granted. Bioconversions aimed at industry-scale vanillin produc- enzyme catalyzing conversion of ferulic acid and its glucoside
tion using microorganisms based on vanillin-like structural com- into vanillin and its glucoside offers new opportunities for the
pounds, such as eugenol, isoeugenol, and ferulic acid, have Vanilla pod–based industries. The regulation and accumulation
been extensively studied in the last two decades. Recently, of vanillin glucoside in the capsules of cultivated vines in
vanillin production from cheaper non-toxic glucose has gained response to abiotic and biotic environmental challenges may
focus. The increasing knowledge of metabolic pathways leading now be assessed at the molecular level. Likewise, in general,
to vanillin production and identification and characterization of each plant species contains about 50 different papain-like
the enzymes and genes involved offer new opportunities for cysteine proteinases (Garcia-Lorenzo et al., 2006). They play a
bioengineering of industrially applicable microorganisms by sta- role in the turnover of proteins at specific time points of plant
ble integration of expression cassettes required for vanillin syn- growth and development, e.g. in a ripening fruit. In the course
thesis. Industrially applicable microorganisms that are able to of evolution of the vanilla-producing V. planifolia orchid, natural
yield more than 1 g/l of vanillin are listed in Table 1. mutations arose in one of these cysteine proteinases, enabling
the enzyme to convert ferulic acid into vanillin. Many orchids
The available microbial bioengineering platforms for vanillin closely related to the vanilla orchid are present in nature but
production have a few general drawbacks, as emphasized in pre- appear not to be able to synthesize vanillin. Classic mutation
vious sections, and optimization of these microbial production breeding to obtain vanillin production in such species may now
systems is necessary to meet the increasing demand for bio- be initiated. This would offer a more diverse production system
logically produced vanillin. Much attention has been given to less prone to diseases, and maybe the introduction of vanillin
cost and yield optimization by searching for cheaper substrates, synthesizing orchids, which would have natural pollinators in their
efficient downstream processing, shorter incubation time, etc., growth habitats and thus not require hand pollination by humans.
with the overall aim of lowering the production cost of bio-
technologically produced vanillin in order to gain a competitive The discovery of the vanillin synthase enzyme has direct implica-
market position in relation to chemically synthesized vanillin. tions for use of vanillin in agricultural industries. Ferulic acid is a
The microbial cell factories so far studied for biotechnology- key intermediate in lignin monomer formation in plants. In contrast
derived vanillin production are mainly based on E. coli strains to the situation in microorganisms, ferulic acid and its glucoside are
and on a few fungal and yeast strains. Nowadays, genome endogenously produced in plant cells and readily available as sub-
sequencing of organisms may be achieved in a cost-efficient strates for vanillin synthase. If so desired, transgenic plants with
manner and the time needed to determine complete microbial high vanillin synthase activity may be used as production sources
genomes has dramatically decreased. Such datasets would for vanillin glucoside. Stable expression of VpVAN and, e.g.
certainly guide targeted improvement of the microorganisms AtUGT72E2 in plants would be expected to result in vanillin gluco-
that have already been studied and would provide an improved side formation in varying amounts. Vanillin is an animal feeding
genetic basis to engineer microorganisms that have not previ- stimulant, and vanillin production in barley grains used for pig
ously been comprehensively studied and thus not considered feed may thus have commercial potential. As demonstrated,
as production hosts. Vanillin production in microorganisms using molasses may be used for vanillin production based on their ferulic
bioengineering would greatly benefit from identification of micro- acid content and following supplementation with yeast expressing
organisms that are more tolerant to vanillin and to the precursors vanillin synthase but are devoid of ferulate decarboxylase activity.
used for its production. In addition, identification of alternative
pathways for vanillin formation that are better suited for genetic The primary economic driving force for the biotechnology-
engineering of production strains is likely to take advantage of derived flavor industry is the desire to establish reliable and
the fast-growing genomic databases. It should also be high- economically profitable production systems that are environ-
lighted that there is a need to exploit extended use of already mentally benign in comparison with the classic production
approved natural GRAS microorganisms for vanillin production; approaches based on large-scale organic chemical synthesis.
Molecular Plant 8, 40–57, January 2015 ª The Author 2015. 53
Molecular Plant Vanillin–Bioconversion and Bioengineering
An additional driver is the fact that flavor compounds produced Received: July 8, 2014
from natural raw materials by microbial or enzymatic methods Accepted: September 15, 2014
in accordance with European and US legislation are labeled Published: September 30, 2014
as ‘‘natural’’. This type of labeling is to the benefit of the manu-
facturer, considering the current consumer trends whereby REFERENCES
products used in the food and flavor sector labeled ‘‘natural’’ Abraham, W.R., Arfmann, H.A., Stumpf, B., Washausen, P., and
are preferred and thus gain a higher sales price. From the Kieslich, K. (1988). Microbial transformations of some terpenoids
and natural compounds. In Bioflavour 0 87, P. Schreier, ed. (Berlin:
point of view of the consumers, the label ‘‘natural’’ may be
Walter de Gruyter), pp. 399–414.
misleading, because a majority of consumers would then be
deluded into attributing the flavor compound to the plant Achterholt, S., Priefert, H., and Steinbuchel, A. (1998). Purification
and characterization of the coniferyl aldehyde dehydrogenase from
species known as the common original source. A proper way
Pseudomonas sp. strain HR199 and molecular characterization of
to signify that a flavor compound known from nature was pro-
the gene. J. Bacteriol. 180:4387–4391.
duced biotechnologically in microorganisms could be to add
the prefix ‘‘bio-‘‘ to the name of the flavor compound to inform Achterholt, S., Priefert, H., and Steinbuchel, A. (2000). Identification of
Amycolatopsis sp strain HR167 genes, involved in the bioconversion of
the consumer that the flavor has been produced using biotech-
ferulic acid to vanillin. Appl. Microbiol. Biotechnol. 54:799–807.
nological approaches. This type of openness would offer many
long-term benefits. It would enable the consumer to clearly Ashengroph, M., Nahvi, I., Zarkesh-Esfahani, H., and Momenbeik, F.
distinguish between a natural extract containing a complex (2010). Optimization of media composition for improving conversion
of isoeugenol into vanillin with Pseudomonas sp. strain KOB10 using
mixture of flavor compounds and a biotechnologically pro-
the Taguchi method. Biocatal. Biotransform. 28:339–347.
duced product composed of one of the most important flavor
compounds of the natural extract. The biotechnologically pro- Ashengroph, M., Nahvi, I., Zarkesh-Esfahani, H., and Momenbeik, F.
duced flavors should thus rather be considered and promoted (2012). Conversion of isoeugenol to vanillin by Psychrobacter sp.
strain CSW4. Appl. Biochem. Biotechnol. 166:1–12.
as environmentally benign substitutes for identical compounds
currently obtained by large-scale chemical synthesis. Chemi- Barghini, P., Di Gioia, D., Fava, F., and Ruzzi, M. (2007). Vanillin
cally synthesized flavor compounds and the introduction of production using metabolically engineered Escherichia coli under
flavors not occurring in nature are in general subject to less non-growing conditions. Microb. Cell Fact. 6.
consumer acceptance. Such products are labeled as ‘‘nature Bauer, K., Garbe, D., and Surburg, H. (2008). Common Fragrance and
identical’’ and ‘‘artificial’’ (EC Flavor Directive 88/388/EEC) Flavor Materials: Preparation, Properties and Uses (Weinheim: John
(US Code of Federal Regulation 21 CFR 101.22). This has Wiley).
reduced the market value of flavors produced by chemical Benen, J.A., Sánchez-Torres, P., Wagemaker, M.J., Fraaije, M.W., van
synthesis. Berkel, W.J., and Visser, J. (1998). Molecular cloning, sequencing, and
heterologous expression of the vaoA gene from Penicillium
The market launch of several biotechnology-derived vanillin prod- simplicissimum CBS 170.90 encoding vanillyl-alcohol oxidase. J. Biol.
Chem. 273:7865–7872.
ucts has generated some media attention (nytimes.com) and
has awakened public suspicion and demand for government Bennett, J.P., Bertin, L., Moulton, B., Fairlamb, I.J., Brzozowski, A.M.,
oversight. It is important to reconsider whether or not it is Walton, N.J., and Grogan, G. (2008). A ternary complex of
justified to label biotechnologically produced flavors such as hydroxycinnamoyl-CoA hydratase-lyase (HCHL) with acetyl-CoA and
vanillin gives insights into substrate specificity and mechanism.
vanillin as ‘‘natural’’. The general public cannot be expected to
Biochem. J. 414:281–289.
understand and adapt to definitions of flavor codes that are not
self-evident and obvious. Unclear labeling policies serve to in- Benoit, I., Navarro, D., Marnet, N., Rakotomanomana, N., Lesage-
crease consumer suspicions toward biotechnological-derived Meessen, L., Sigoillot, J.C., Asther, M., and Asther, M. (2006).
Feruloyl esterases as a tool for the release of phenolic compounds
flavors and toward the entire food industry. The three main flavor
from agro-industrial by-products. Carbohydr. Res. 341:1820–1827.
codes ‘‘natural’’, ‘‘nature identical flavor,’’ and ‘‘artificial’’ do not
properly specify and distinguish commercially available products. Berg J.M., Tymoczko J.L., and Stryer L., eds. (2012). Biochemistry
This situation obviously results in confused consumers. Organi- (New York: W.H. Freeman), pp. 670–695.
zations like Friends of the Earth and Greenpeace profit from this Berger, R.G. (2007). Flavours and Fragrances: Chemistry, Bioprocessing
unclear situation by establishing a communication platform to and Sustainability (Berlin: Springer).
voice their general resistance to all products obtained using Bhathena, J., Kulamarva, A., Urbanska, A.M., Martoni, C., and
genetic engineering, even when no genetic material is present Prakash, S. (2007). Microencapsulated bacterial cells can be used
in the commercialized product. to produce the enzyme feruloyl esterase: preparation and in-vitro
analysis. Appl. Microbiol. Biotechnol. 75:1023–1029.
In conclusion, biotechnological production of vanillin using safe Bonnin, E., Grange, H., Lesage-Meessen, L., Asther, M., and Thibault,
substrate precursors, food-grade production organisms, and J.F. (2000). Enzymic release of cellobiose from sugar beet pulp, and its
environmentally benign and economically feasible downstream use to favour vanillin production in Pycnoporus cinnabarinus from
processing is envisioned to result in a compatible and sustainable vanillic acid. Carbohydr. Polym. 41:143–151.
alternative to vanillin produced by chemical synthesis. Brochado, A.R., Matos, C., Møller, B.L., Hansen, J., Mortensen, U.H.,
and Patil, K.R. (2010). Improved vanillin production in baker’s yeast
through in silico design. Microb. Cell Fact. 9:84.
ACKNOWLEDGMENTS
The authors have filed a patent application on the VpVAN. Birger Lindberg Brunel, F., and Davison, J. (1988). Cloning and sequencing of
Møller is a member of the Science Advisory Board of Evolva, a company Pseudomonas genes encoding vanillate demethylase. J. Bacteriol.
who has vested interests in vanillin. 170:4924–4930.

54 Molecular Plant 8, 40–57, January 2015 ª The Author 2015.


Vanillin–Bioconversion and Bioengineering Molecular Plant
Cambra, I., Hernandez, D., Diaz, I., and Martinez, M. (2012). Structural Garcia-Lorenzo, M., Sjodin, A., Jansson, S., and Funk, C. (2006).
basis for specificity of propeptide-enzyme interaction in barley C1A Protease gene families in Populus and Arabidopsis. BMC Plant Biol. 6:30.
cysteine peptidases. PLoS One 7:e37234.
Gasson, M.J., Kitamura, Y., McLauchlan, W.R., Narbad, A., Parr, A.J.,
Cao, X.H., Cui, Y.Q., Liao, X.H., Zhou, G.T., and Thamm, L. (2010). Gene Parsons, E.L., Payne, J., Rhodes, M.J., and Walton, N.J. (1998).
cloning of phenolic acid decarboxylase from Bacillus subtilis and Metabolism of ferulic acid to vanillin–a bacterial gene of the enoyl-
expression in top-fermenting yeast strain. Afr. J. Biotechnol. 9:5284– SCoA hydratase/isomerase superfamily encodes an enzyme for the
5291. hydration and cleavage of a hydroxycinnamic acid SCoA thioester.
J. Biol. Chem. 273:4163–4170.
Casey, J., and Dobb, R. (1992). Microbial routes to aromatic-aldehydes.
Enzyme Microb. Technol. 14:739–747. Gross, G.G., and Zenk, M.H. (1969). Reduction of aromatic acids to
aldehydes and alcohols in the cell-free system. 2. Purification and
Clark, G.S. (1990). Vanillin: a profile. Perfum. Flavor. 15:45–54.
properties of aryl-alcohol: NADP-oxidoreductase from Neurospora
deVries, R.P., Michelsen, B., Poulsen, C.H., Kroon, P.A., van den crassa. Eur. J. Biochem. 8:420–425.
Heuvel, R.H., Faulds, C.B., Williamson, G., van den Hombergh,
Gurujeyalakshmi, G., and Mahadevan, A. (1987). Dissimilation of ferulic
J.P., and Visser, J. (1997). The faeA genes from Aspergillus niger
acid by Bacillus subtilis. Curr. Microbiol. 16:69–73.
and Aspergillus tubingensis encode ferulic acid esterases involved in
degradation of complex cell wall polysaccharides. Appl. Environ. Hansen, E.H., Møller, B.L., Kock, G.R., Bünner, C.M., Kristensen, C.,
Microbiol. 63:4638–4644. Jensen, O.R., Okkels, F.T., Olsen, C.E., Motawia, M.S., and
Hansen, J. (2009). De novo biosynthesis of vanillin in fission yeast
Dewick, P.M. (1989). The biosynthesis of shikimate metabolites. Nat.
(Schizosaccharomyces pombe) and baker’s yeast (Saccharomyces
Prod. Rep. 6:263–290.
cerevisiae). Appl. Environ. Microbiol. 75:2765–2774.
Di Gioia, D., Luziatelli, F., Negroni, A., Ficca, A.G., Fava, F., and Ruzzi, M.
Hansen, J., Hansen, E.H., Sompalli, H.P., Sheridan, J.M., Heal, J.R.
(2011). Metabolic engineering of Pseudomonas fluorescens for the
and Hamilton, W.D.O. (2013). WO2013022881 A1, World Intellectual
production of vanillin from ferulic acid. J. Biotechnol. 156:309–316.
Property Organization, USA.
Dignum, M.J.W., Kerler, J., and Verpoorte, R. (2001). Vanilla production:
Harris, P.J., and Trethewey, J.A.K. (2010). The distribution of ester-linked
technological, chemical, and biosynthetic aspects. Food Rev. Int.
ferulic acid in the cell walls of angiosperms. Phytochem. Rev. 9:19–33.
17:199–219.
Higuchi, T., Ito, Y., and Kawamura, I. (1967). p-Hydroxyphenylpropane
Faulds, C.B., Bartolome, B., and Williamson, G. (1997). Novel
component of grass lignin and role of tyrosine-ammonia lyase in its
biotransformations of agro-industrial cereal waste by ferulic acid
formation. Phytochemistry 6:875.
esterases. Ind. Crops Prod. 6:367–374.
Hocking, M.B. (1997). Vanillin: synthetic flavoring from spent sulfite liquor.
Ferulic acid ex Rice bran, Novorate, 2014. http://www.novorate.com. 01
J. Chem. Educ. 74:1055–1059.
August, 2014.
Hua, D., Ma, C., Song, L., Lin, S., Zhang, Z., Deng, Z., and Xu, P.
Fleige, C., Hansen, G., Kroll, J., and Steinbuchel, A. (2013).
(2007a). Enhanced vanillin production from ferulic acid using
Investigation of the Amycolatopsis sp strain ATCC 39116 vanillin
adsorbent resin. Appl. Microbiol. Biotechnol. 74:783–790.
dehydrogenase and its impact on the biotechnical production of
vanillin. Appl. Environ. Microbiol. 79:81–90. Hua, D., Ma, C., Lin, S., Song, L., Deng, Z., Maomy, Z., Zhang, Z.,
Yu, B., and Xu, P. (2007b). Biotransformation of isoeugenol to
Food Additive Status List, U.S. Food and Drug Administration. 01
vanillin by a newly isolated Bacillus pumilus strain: identification of
August 2014. http://www.fda.gov/food/ingredientspackaginglabeling/
major metabolites. J. Biotechnol. 130:463–470.
foodadditivesingredients/ucm091048.htm.
Huang, Z.X., Dostal, L., and Rosazza, J.P.N. (1993). Microbial
Fritz, R.R., Hodgins, D.S., and Abell, C.W. (1976). Phenylalanine
transformations of ferulic acid by Saccharomyces cerevisiae and
ammonia-lyase. Induction and purification from yeast and clearance
Pseudomonas fluorescens. Appl. Environ. Microbiol. 59:2244–2250.
in mammals. J. Biol. Chem. 251:4646–4650.
Huang, Z.X., Dostal, L., and Rosazza, J.P.N. (1994). Purification and
Fu, T.J., Singh, G., and Curtis, W.R. (1999). Plant Cell and Tissue
characterization of a ferulic acid decarboxylase from Pseudomonas
Culture for the Production of Food Ingredients (New York: Springer),
fluorescens. J. Bacteriol. 176:5912–5918.
pp. 237–250.
Jensen, K.A., Evans, K.M.C., Kirk, T.K., and Hammel, K.E. (1994).
Funk, C., and Brodelius, P.E. (1990a). Phenylpropanoid metabolism
Biosynthetic pathway for veratryl alcohol in the ligninolytic fungus
in suspension-cultures of Vanilla planifolia.1. Influence of growth-
Phanerochaete chrysosporium. Appl. Environ. Microbiol. 60:709–714.
regulators and an elicitor on phenylpropanoid metabolism in
suspension-cultures of Vanilla planifolia. Phytochemistry 29:845–848. Jin, J., Mazon, H., van den Heuvel, R.H., Janssen, D.B., and Fraaije,
M.W. (2007). Discovery of a eugenol oxidase from Rhodococcus sp.
Funk, C., and Brodelius, P.E. (1990b). Phenylpropanoid metabolism in
strain RHA1. FEBS J. 274:2311–2321.
suspension-cultures of Vanilla planifolia Andr. 2. Effects of precursor
feeding and metabolic-inhibitors. Plant Physiol. 94:95–101. Johnson, T.S., Ravishankar, G.A., and Venkataraman, L.V. (1996).
Biotransformation of ferulic acid and vanillylamine to capsaicin and
Furukawa, H., Wieser, M., Morita, H., and Nagasawa, T. (1999).
vanillin in immobilized cell cultures of Capsicum frutescens. Plant
Microbial synthesis of coniferyl alcohol by the fungus Byssochlamys
Cell Tissue Organ Cult. 44:117–121.
fulva V107. Biosci. Biotechnol. Biochem. 63:1141–1142.
Jurkova, M., and Wurst, M. (1993). Biodegradation of aromatic
Gabriac, B., Werck-Reichhart, D., Teutsch, H., and Durst, F. (1991).
carboxylic-acids by Pseudomonas mira. FEMS Microbiol. Lett.
Purification and immunocharacterization of a plant cytochrome
111:245–250.
P450: the cinnamic acid 4-hydroxylase. Arch. Biochem. Biophys.
288:302–309. Karapinar, M. (1990). Inhibitory effects of anethole and eugenol on the
growth and toxin production of Aspergillus parasiticus. Int. J. Food.
Gallage, N.J., Hansen, E.H., Kannangara, R., Olsen, C.E., Motawia,
Microbiol. 10:193–199.
M.S., Jørgensen, K., Holme, I., Hebelstrup, K., Grisoni, M., and
Møller, B.L. (2014). Vanillin formation from ferulic acid in Vanilla Karmakar, B., Vohra, R.M., Nandanwar, H., Sharma, P., Gupta, K.G.,
planifolia is catalysed by a single enzyme. Nat. Commun. 5. and Sobti, R.C. (2000). Rapid degradation of ferulic acid via

Molecular Plant 8, 40–57, January 2015 ª The Author 2015. 55


Molecular Plant Vanillin–Bioconversion and Bioengineering
4-vinylguaiacol and vanillin by a newly isolated strain of Bacillus Martinez-Cuesta, M.D., Payne, J., Hanniffy, S.B., Gasson, M.J., and
coagulans. J. Biotechnol. 80:195–202. Narbad, A. (2005). Functional analysis of the vanillin pathway in a
Kasana, R.C., Sharma, U.K., Sharma, N., and Sinha, A.K. (2007). vdh-negative mutant strain of Pseudomonas fluorescens AN103.
Isolation and identification of a novel strain of Pseudomonas Enzyme Microb. Technol. 37:131–138.
chlororaphis capable of transforming isoeugenol to vanillin. Curr. Masai, E., Harada, K., Peng, X., Kitayama, H., Katayama, Y., and
Microbiol. 54:457–461. Fukuda, M. (2002). Cloning and characterization of the ferulic
Kirk, T.K., and Farrell, R.L. (1987). Enzymatic combustion–the microbial- acid catabolic genes of Sphingomonas paucimobilis SYK-6. Appl.
degradation of lignin. Annu. Rev. Microbiol. 41:465–505. Environ. Microbiol. 68:4416–4424.

Knorr, K., Beaumont, M.D., Caster, C.S., Doe ^ rnenburg, H., Gross, B., Mitra, A., Kitamura, Y., Gasson, M.J., Narbad, A., Parr, A.J., Payne, J.,
Pandya, Y., and Romagnoli, L.G. (1990). Plant-tissue culture for the Rhodes, M.J., Sewter, C., and Walton, N.J. (1999). 4-
production of naturally derived food ingredients. Food Technol. 44:71–79. Hydroxycinnamoyl-CoA hydratase/lyase (HCHL)–an enzyme of
phenylpropanoid chain cleavage from Pseudomonas. Arch.
Koeduka, T., Fridman, E., Gang, D.R., Vassão, D.G., Jackson, B.L., Biochem. Biophys. 365:10–16.
Kish, C.M., Orlova, I., Spassova, S.M., Lewis, N.G., Noel, J.P.,
et al. (2006). Eugenol and isoeugenol, characteristic aromatic Muheim, A., and Lerch, K. (1999). Towards a high-yield bioconversion of
constituents of spices, are biosynthesized via reduction of a coniferyl ferulic acid to vanillin. Appl. Microbiol. Biotechnol. 51:456–461.
alcohol ester. Proc. Natl. Acad. Sci. U S A 103:10128–10133. Mukai, N., Masaki, K., Fujii, T., Kawamukai, M., and Iefuji, H. (2010).
Krings, U., and Berger, R.G. (1998). Biotechnological production of PAD1 and FDC1 are essential for the decarboxylation of phenylacrylic
flavours and fragrances. Appl. Microbiol. Biotechnol. 49:1–8. acids in Saccharomyces cerevisiae. J. Biosci. Bioeng. 109:564–569.

Krings, U., Hinz, M., and Berger, R.G. (1996). Degradation of [H-2] Narbad, A., and Gasson, M.J. (1998). Metabolism of ferulic acid via
phenylalanine by the basidiomycete Ischnoderma benzoinum. J. vanillin using a novel CoA-dependent pathway in a newly-isolated
Biotechnol. 51:123–129. strain of Pseudomonas fluorescens. Microbiology-UK 144:1397–1405.

Kunc, F., and Macura, J. (1966). Decomposition of root exudates in soil. Nazareth, S., and Mavinkurve, S. (1986). Degradation of ferulic acid via
Folia Microbiol. 11:239. 4-vinylguaiacol by Fusarium solani (Mart) Sacc. Can. J. Microbiol.
32:494–497.
Lam, K.C., Ibrahim, R.K., Behdad, B., and Dayanandan, S. (2007).
Structure, function, and evolution of plant O-methyltransferases. Okeke, B.C., and Venturi, V. (1999). Construction of recombinants
Genome 50:1001–1013. Pseudomonas putida BO14 and Escherichia coli QEFCA8 for ferulic
acid biotransformation to vanillin. J. Biosci. Bioeng. 88:103–106.
Lambert, F., Zucca, J., Ness, F., and Aigle, M. (2014). Production of
ferulic acid and coniferyl alcohol by conversion of eugenol using a Otto, H.H., and Schirmeister, T. (1997). Cysteine proteases and their
recombinant strain of Saccharomyces cerevisiae. Flavour Fragrance inhibitors. Chem. Rev. 97:133–171.
J. 29:14–21. Overhage, J., Priefert, H., Rabenhorst, J., and Steinbuchel, A. (1999).
Lee, E.G., Yoon, S.H., Das, A., Lee, S.H., Li, C., Kim, J.Y., Choi, M.S., Biotransformation of eugenol to vanillin by a mutant of Pseudomonas
Oh, D.K., and Kim, S.W. (2009). Directing vanillin production from sp strain HR199 constructed by disruption of the vanillin
ferulic acid by increasedacetyl-CoA consumption in recombinant dehydrogenase (vdh) gene. Appl. Microbiol. Biotechnol. 52:820–828.
Escherichia coli. Biotechnology and Bioengineering 102:200–208. Overhage, J., Priefert, H., Rabenhorst, J., and Steinbuechel, A. (2000).
Lesage-Meessen, L., Delattre, M., Haon, M., Thibault, J.F., Ceccaldi, Google Patents. Patent number DE19850242-A1/30, 04-MAY-2000.
B.C., Brunerie, P., and Asther, M. (1996). A two-step bioconversion HAARMANN & REIMER GMBH, Germany.
process for vanillin production from ferulic acid combining Aspergillus Overhage, J., Steinbuchel, A., and Priefert, H. (2003). Highly efficient
niger and Pycnoporus cinnabarinus. J. Biotechnol. 50:107–113. biotransformation of eugenol to ferulic acid and further conversion to
Lesage-Meessen, L., Stentelaire, C., Lomascolo, A., Couteau, D., vanillin in recombinant strains of Escherichia coli. Appl. Environ.
Asther, M., Moukha, S., Record, E., Sigoillot, J.C., and Asther, M. Microbiol. 69:6569–6576.
(1999). Fungal transformation of ferulic acid from sugar beet pulp to Overhage, J., Steinbuchel, A., and Priefert, H. (2006). Harnessing eugenol
natural vanillin. J. Sci. Food Agric. 79:487–490. as a substrate for production of aromatic compounds with recombinant
Lesage-Meessen, L., Lomascolo, A., Bonnin, E., Thibault, J.F., strains of Amycolatopsis sp HR167. J. Biotechnol. 125:369–376.
Buleon, A., Roller, M., Asther, M., Record, E., Ceccaldi, B.C., and Plaggenborg, R., Steinbuchel, A., and Priefert, H. (2001). The
Asther, M. (2002). A biotechnological process involving filamentous coenzyme A-dependent, non-beta-oxidation pathway and not direct
fungi to produce natural crystalline vanillin from maize bran. Appl. deacetylation is the major route for ferulic acid degradation in Delftia
Biochem. Biotechnol. 102:141–153. acidovorans. FEMS Microbiol. Lett. 205:9–16.
Li, K., and Frost, J.W. (1998). Synthesis of vanillin from glucose. J. Am. Plaggenborg, R., Overhage, J., Steinbuchel, A., and Priefert, H. (2003).
Chem. Soc. 120:10545–10546. Functional analyses of genes involved in the metabolism of ferulic
Li, X.M., Yang, J., Li, X., Gu, W., Huang, J., and Zhang, K.Q. (2008). acid in Pseudomonas putida KT2440. Appl. Microbiol. Biotechnol.
The metabolism of ferulic acid via 4-vinylguaiacol to vanillin by 61:528–535.
Enterobacter sp Px6-4 isolated from Vanilla root. Process Biochem. Plaggenborg, R., Overhage, J., Loos, A., Archer, J.A., Lessard, P.,
43:1132–1137. Sinskey, A.J., Steinbüchel, A., and Priefert, H. (2006). Potential of
Martinez, D., Larrondo, L.F., Putnam, N., Gelpke, M.D., Huang, K., Rhodococcus strains for biotechnological vanillin production from
Chapman, J., Helfenbein, K.G., Ramaiya, P., Detter, J.C., and ferulic acid and eugenol. Appl. Environ. Microbiol. 72:745–755.
Larimer, F. (2004). Genome sequence of the lignocellulose Pollack, A., What’s That Smell? Exotic Scents Made From Re-engineered
degrading fungus Phanerochaete chrysosporium strain RP78 (vol 22, Yeast. The New York Times. 21 October, 2013. http://www.nytimes.
pg 695, 2004). Nat. Biotechnol. 22:899. com/2013/10/21/business/whats-that-smell-exotic-scents-made-from-
Martinez-Cuesta, M.C., Gasson, M.J., and Narbad, A. (2005). re-engineered-yeast.html?pagewanted=all%_r=0. 01 Aug 2014.
Heterologous expression of the plant coumarate: CoA ligase in Priefert, H., Rabenhorst, J., and Steinbuchel, A. (1997). Molecular
Lactococcus lactis. Lett. Appl. Microbiol. 40:44–49. characterization of genes of Pseudomonas sp. strain HR199 involved

56 Molecular Plant 8, 40–57, January 2015 ª The Author 2015.


Vanillin–Bioconversion and Bioengineering Molecular Plant
in bioconversion of vanillin to protocatechuate. J. Bacteriol. 179:2595– Tadasa, K. (1977). Degradation of eugenol by a microorganism. Agric.
2607. Biol. Chem. 41:925–929.
Priefert, H., Overhage, J., and Steinbuchel, A. (1999). Identification The flavor that carries - Vanillin for 50 years, Borregaard. 19 April, 2012.
and molecular characterization of the eugenol hydroxylase genes http://www.borregaard.com/News/The-flavor-that-carries-Vanillin-for-
(ehyA/ehyB) of Pseudomonas sp. strain HR199. Arch. Microbiol. 50-years/(language)/eng-GB. 01 August, 2014.
172:354–363. Toms, A., and Wood, J.M. (1970). The degradation of trans-ferulic acid by
Priefert, H., Rabenhorst, J., and Steinbuchel, A. (2001). Pseudomonas acidovorans. Biochemistry 9.
Biotechnological production of vanillin. Appl. Microbiol. Biotechnol. Topakas, E., Kalogeris, E., Kekos, D., Macris, B.J., and
56:296–314. Christakopoulos, P. (2003). Bioconversion of ferulic acid into vanillic
Rabenhorst, J. (1996). Production of methoxyphenol-type natural aroma acid by the thermophilic fungus Sporotrichum thermophile. LWT –
chemicals by biotransformation of eugenol with a new Pseudomonas Food Sci. Technol. 36:561–565.
sp. Appl. Microbiol. Biotechnol. 46:470–474. Torres, B.R., Aliakbarian, B., Torre, P., Perego, P., Domı́nguez, J.M.,
Rahouti, M., Seiglemurandi, F., Steiman, R., and Eriksson, K.E. (1989). Zilli, M., and Converti, A. (2009). Vanillin bioproduction from alkaline
Metabolism of ferulic acid by Paecilomyces variotii and Pestalotia hydrolyzate of corn cob by Escherichia coli JM109/pBB1. Enzyme
palmarum. Appl. Environ. Microbiol. 55:2391–2398. Microb. Technol. 44:154–158.
Rao, S.R., and Ravishankar, G.A. (2000). Biotransformation of Turk, V., Stoka, V., Vasiljeva, O., Renko, M., Sun, T., Turk, B., and Turk,
protocatechuic aldehyde and caffeic acid to vanillin and capsaicin D. (2012). Cysteine cathepsins: from structure, function and regulation
in freely suspended and immobilized cell cultures of Capsicum to new frontiers. Biochim. Biophys. Acta 1824:68–88.
frutescens. J. Biotechnol. 76:137–146. Vanilla. (2014). A sustainable production route. Evolva A/S. www.evolva/
Rao, S.R., and Ravishankar, G.A. (2000). Vanilla flavour: production by products/vanilla, 01 August, 2014.
conventional and biotechnological routes. J. Sci. Food Agric. Vanillin Natural (ex-ferulic acid), De Monchy Aromatics, 2014. http://
80:289–304. www.demonchyaromatics.com/en/products/vanillin-natural-(ex-ferulic-
Rhovanil Natural, Solvay. 2014. http://www.safevanillin.com/en/vanillin- acid)/. 01 August, 2014.
and-ethyl-vanillin-range/rhovanil-natural/index.html. 01 August, 2014. Venkitasubramanian, P., Daniels, L., and Rosazza, J.P. (2007). Reduction
Rosazza, J.P.N., Huang, Z., Dostal, L., Volm, T., and Rousseau, B. of carboxylic acids by Nocardia aldehyde oxidoreductase requires a
(1995). Review: biocatalytic transformations of ferulic acid: an phosphopantetheinylated enzyme. J. Biol. Chem. 282:478–485.
abundant aromatic natural product. J. Ind. Microbiol. 15:457–471. Venturi, V., Zennaro, F., Degrassi, G., Okeke, B.C., and Bruschi, C.V.
Russell, D.W., and Conn, E.E. (1967). The cinnamic acid 4-hydraxylase of (1998). Genetics of ferulic acid bioconversion to protocatechuic
pea seedlings. Arch. Biochem. Biophys. 122:256–258. acid in plant-growth-promoting Pseudomonas putida WCS358.
Ruzzi, M., Barghini, P., Montebove, F., and Ponente, A.S. (1997). Effect of Microbiology-UK 144:965–973.
the carbon source on the utilization of ferulic, m- and p-coumaric acids by Walton, N.J., Narbad, A., Faulds, C.B., and Williamson, G. (2000). Novel
a Pseudomonas fluorescens strain. Ann. Microbiol. Enzimol. 47:87–96. approaches to the biosynthesis of vanillin. Curr. Opin. Biotechnol.
Ryu, J.Y., Seo, J., Ahn, J.H., Sadowsky, M.J., and Hur, H.G. (2012). 11:490–496.
Transcriptional control of the isoeugenol monooxygenase of Walton, N.J., Mayer, M.J., and Narbad, A. (2003). Vanillin.
Pseudomonas nitroreducens Jin1 in Escherichia coli. Biosci. Phytochemistry 63:505–515.
Biotechnol. Biochem. 76:1891–1896. Yamada, M., Okada, Y., Yoshida, T., and Nagasawa, T. (2007).
Sabisch, M., and Smith, D. (2014). The Complex Regulatory Landscape Biotransformation of isoeugenol to vanillin by Pseudomonas putida
for Natural Flavor Ingredients. Sigma Aldrich. http://www.sigmaaldrich. IE27 cells. Appl. Microbiol. Biotechnol. 73:1025–1030.
com, 01 August, 2014. Yamada, M., Okada, Y., Yoshida, T., and Nagasawa, T. (2008). Vanillin
Schoch, G., Goepfert, S., Morant, M., Hehn, A., Meyer, D., Ullmann, P., production using Escherichia coli cells over-expressing isoeugenol
and Werck-Reichhart, D. (2001). CYP98A3 from Arabidopsis thaliana monooxygenase of Pseudomonas putida. Biotechnol. Lett. 30:665–670.
is a 30 -hydroxylase of phenolic esters, a missing link in the Yang, W.W., Tang, H., Ni, J., Wu, Q., Hua, D., Tao, F., and Xu, P. (2013).
phenylpropanoid pathway. J. Biol. Chem. 276:36566–36574. Characterization of two Streptomyces enzymes that convert ferulic
Sense Capture Vanillin by Mane (2014). Mane. http://www.mane.com/ acid to vanillin. PLoS One 8.
sites/default/files/file-block/sense_capture_vanillin.pdf, 01 August, 2014. Yoon, S.H., Li, C., Lee, Y.M., Lee, S.H., Kim, S.H., Choi, M.S., Seo, W.T.,
Shimoni, E., Ravid, U., and Shoham, Y. (2000). Isolation of a Bacillus sp Yang, J.K., Kim, J.Y., and Kim, S.W. (2005). Production of vanillin
capable of transforming isoeugenol to vanillin. J. Biotechnol. 78:1–9. from ferulic acid using recombinant strains of Escherichia coli.
Shimoni, E., Baasov, T., Ravid, U., and Shoham, Y. (2002). The trans- Biotechnol. Bioprocess Eng. 10:378–384.
anethole degradation pathway in an Arthrobacter sp. J. Biol. Chem. Zenk, M.H., Ulbrich, B., Busse, J., and Stockigt, J. (1980). Procedure for
277:11866–11872. the enzymatic-synthesis and isolation of cinnamoyl-CoA thiolesters
Sinha, A.K., Sharma, U.K., and Sharma, N. (2008). A comprehensive using a bacterial system. Anal. Biochem. 101:182–187.
review on vanilla flavor: extraction, isolation and quantification of Zhang, Y.M., Xu, P., Han, S., Yan, H.Q., and Ma, C.Q. (2006).
vanillin and others constituents. Int. J. Food. Sci. Nutr. 59:299–326. Metabolism of isoeugenol via isoeugenol-diol by a newly isolated
Stentelaire, C., Lesage-Meessen, L., Delattre, M., Haon, M., Sigoillot, strain of Bacillus subtilis HS8. Appl. Microbiol. Biotechnol. 73:771–779.
J.C., Ceccaldi, B.C., and Asther, M. (1998). By-passing of unwanted Zhao, L.Q., Sun, Z.H., Zheng, P.Z., and Lei, L. (2005). Biotransformation
vanillyl alcohol formation using selective adsorbents to improve vanillin of isoeugenol to vanillin by a novel strain of Bacillus fusiformis.
production with Phanerochaete chrysosporium. World J. Microbiol. Biotechnol. Lett. 27:1505–1509.
Biotechnol. 14:285–287. Zheng, L.R., Zheng, P., Sun, Z., Bai, Y., Wang, J., and Guo, X. (2007).
Sutherland, J.B., Crawford, D.L., and Pometto, A.L. (1983). Metabolism Production of vanillin from waste residue of rice bran oil by
of cinnamic, p-coumaric, and ferulic acids by Streptomyces setonii. Aspergillus niger and Pycnoporus cinnabarinus. Bioresour. Technol.
Can. J. Microbiol. 29:1253–1257. 98:1115–1119.

Molecular Plant 8, 40–57, January 2015 ª The Author 2015. 57

You might also like