You are on page 1of 4

Journal of Biotechnology 243 (2017) 25–28

Contents lists available at ScienceDirect

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

Short communication

Biotechnological production of vanillin using immobilized enzymes


Toshiki Furuya a,b,∗ , Mari Kuroiwa a , Kuniki Kino a,∗
a
Department of Applied Chemistry, Faculty of Science and Engineering, Waseda University, 3-4-1 Ohkubo, Shinjuku-ku, Tokyo 169-8555, Japan
b
Department of Applied Biological Science, Faculty of Science and Technology, Tokyo University of Science, 2641 Yamazaki, Noda, Chiba 278-8510, Japan

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

Article history: Vanillin is an important and popular plant flavor, but the amount of this compound available from plant
Received 1 November 2016 sources is very limited. Biotechnological methods have high potential for vanillin production as an alter-
Received in revised form native to extraction from plant sources. Here, we report a new approach using immobilized enzymes for
26 December 2016
the production of vanillin. The recently discovered oxygenase Cso2 has coenzyme-independent catalytic
Accepted 28 December 2016
activity for the conversion of isoeugenol and 4-vinylguaiacol to vanillin. Immobilization of Cso2 on Sep-
Available online 29 December 2016
abeads EC-EA anion-exchange carrier conferred enhanced operational stability enabling repetitive use.
This immobilized Cso2 catalyst allowed 6.8 mg yield of vanillin from isoeugenol through ten reaction
Keywords:
Biocatalysis
cycles at a 1 mL scale. The coenzyme-independent decarboxylase Fdc, which has catalytic activity for
Cascade reaction the conversion of ferulic acid to 4-vinylguaiacol, was also immobilized on Sepabeads EC-EA. We demon-
Ferulic acid strated that the immobilized Fdc and Cso2 enabled the cascade synthesis of vanillin from ferulic acid via
Immobilized enzyme 4-vinylguaiacol with repetitive use of the catalysts. This study is the first example of biotechnological
Vanillin production of vanillin using immobilized enzymes, a process that provides new possibilities for vanillin
production.
© 2016 Elsevier B.V. All rights reserved.

Vanillin (4-hydroxy-3-methoxybenzaldehyde) is the major production (Priefert et al., 2001; Hua et al., 2007; Ma and Daugulis,
aroma component of vanilla, giving it a sweet and creamy odor. 2014).
Its flavor and fragrance properties have made this compound In addition to microorganisms, use of immobilized enzymes
important around the world, especially to the foods and cosmet- shows promise in methods used for vanillin production. Enzymes
ics industries (Rao and Ravishankar, 2000; Walton et al., 2003). in microbial cells and cell-free enzymes are often unstable when
Vanillin is available from the beans of the vanilla orchid through used as biocatalysts. In contrast, immobilized enzymes may be
extraction. However, the yield of vanillin is very low because it more operationally stable due to rigidification of their structure via
accumulates at low levels in the plant (Rao and Ravishankar, 2000; direct interaction with proper carriers. Furthermore, immobilized
Walton et al., 2003). Biotechnological production of vanillin using enzymes can be easily separated from reaction solutions, allowing
microorganisms has attracted much attention as an alternative to repetitive, cost-effective use (Brady and Jordaan, 2009; Hilterhaus
extraction from vanilla beans (Gallage and Møller, 2015; Kaur and et al., 2008; Sheldon and van Pelt, 2013). Although vanillin pro-
Chakraborty, 2013; Priefert et al., 2001). For example, vanillin can duction using microorganisms has been extensively studied, there
be produced from isoeugenol, which occurs in some essential oils, have been no reports concerning production using immobilized
by microorganisms (Priefert et al., 2001; Wangrangsimagul et al., enzymes. This is partly due to the complexity of the enzyme system
2012; Yamada et al., 2007). Ferulic acid is a more practical starting in microorganisms that convert ferulic acid to vanillin. A well-
material for the biotechnological production of vanillin, because known pathway in microorganisms consists of the conversion of
a large amount of this compound can be recovered from agroin- ferulic acid to feruloyl-CoA, catalyzed by feruloyl-CoA synthetase,
dustrial wastes including wheat and rice bran (Di Gioia et al., 2007; followed by the conversion of feruloyl-CoA to vanillin by enoyl-CoA
Rosazza et al., 1995). Many microorganisms that convert ferulic acid hydratase/aldolase; the first step requires ATP and CoA as coen-
to vanillin have been isolated and extensively studied for vanillin zymes (Achterholt et al., 2000; Barghini et al., 2007; Lee et al., 2009;
Narbad and Gasson, 1998). These expensive coenzymes have to be
supplied to continue the coupled enzymatic reactions when the
synthetase and hydratase/aldolase are used as immobilized cata-
lysts.
∗ Corresponding author.
E-mail addresses: tfuruya@rs.tus.ac.jp (T. Furuya), kkino@waseda.jp (K. Kino).

http://dx.doi.org/10.1016/j.jbiotec.2016.12.021
0168-1656/© 2016 Elsevier B.V. All rights reserved.
26 T. Furuya et al. / Journal of Biotechnology 243 (2017) 25–28

(A) 10 8
O2 CH3CHO O

8
6
OCH3 Oxygenase OCH3

of vanillin (mg)
Vanillin (mM)

Total amount
OH Cso2 OH 6
Isoeugenol Vanillin
4
(B) O OH 4
CO2 O2 HCHO O
2
2

OCH3 Decarboxylase OCH3 Oxygenase OCH3


OH Fdc OH Cso2 OH 0 0
Ferulic acid 4-Vinylguaiacol Vanillin 1 2 3 4 5 6 7 8 9 10
Cycle number
Fig. 1. Coenzyme-independent synthesis of vanillin. Vanillin is synthesized from
isoeugenol by oxygenase Cso2 (A) and from ferulic acid via 4-vinylguaiacol by decar- Fig. 2. The vanillin production from isoeugenol by repetitive use of immobilized
boxylase Fdc and oxygenase Cso2 (B). Cso2. Isoeugenol (10 mM) was incubated with 50 mg whole cells or 7 mg protein
immobilized on 100 mg of the carrier in the reaction mixture (1 mL) for 24 h in each
cycle. The amount of protein (7 mg) was almost equal to that obtained through
Recently, we developed a route to vanillin from ferulic acid that disruption of 50 mg whole cells. Vanillin produced in each cycle by the whole cells
does not require any coenzymes (Fig. 1). This artificial pathway con- (gray bars) and the immobilized enzyme (white bars) is shown. Total amount of
sists of a coenzyme-independent decarboxylase (Fdc) that converts vanillin produced by the whole cells (gray circles) and the immobilized enzyme
(white circles) is also shown. Data are the average of three independent experiments,
ferulic acid to 4-vinylguaiacol (2-methoxy-4-vinylphenol), and a
and error bars indicate standard deviation of the mean.
subsequent step with a coenzyme-independent oxygenase (Cso2)
that converts 4-vinylguaiacol to vanillin (Furuya et al., 2014; Furuya
et al., 2015; Muschiol et al., 2015). The ferulic acid decarboxylase led to lower yield of vanillin by the immobilized enzyme than
Fdc from Bacillus pumilus catalyzes the nonoxidative decarboxyla- the whole cells. After the first reaction cycle, the whole cells and
tion of aromatic carboxylic acids (Barthelmebs et al., 2001; Yang the immobilized enzyme were collected from the reaction mixture
et al., 2009). The 4-vinylguaiacol oxygenase Cso2 from Caulobac- and added to a fresh reaction mixture (Fig. 2). In the second reac-
ter segnis belongs to the carotenoid cleavage oxygenase family tion cycle, the whole cells produced only 4.9 mM vanillin. After the
and catalyzes the oxidative cleavage of a conjugated C C bond third reaction cycle, the activity of the whole cells was almost lost.
(Furuya et al., 2014; Kloer and Schulz, 2006). We demonstrated Although the oxygenase Cso2 exhibits high activity in the reac-
that Escherichia coli cells expressing Fdc and Cso2 produced vanillin tion pH 9.0, the enzyme is relatively unstable under the alkaline
from ferulic acid via 4-vinylguaiacol (Furuya et al., 2014). Cso2 condition (Furuya et al., 2014). However, the immobilized enzyme
also converts isoeugenol to vanillin in a coenzyme-independent enabled to produce 6.3 mM vanillin during the second reaction
manner (Furuya et al., 2014) (Fig. 1). This coenzyme-independent cycle. Furthermore, more than 50% of the activity was maintained
feature would be advantageous to the production of vanillin even in the seventh reaction cycle. These results indicated that
using immobilized enzymes. In the present study, we investigated the stability of Cso2 was significantly enhanced by immobilizing
biotechnological production of vanillin using immobilized Fdc and on Sepabeads EC-EA. The total amount of vanillin produced from
Cso2. isoeugenol by the immobilized Cso2 catalyst reached 6.8 mg at the
A proper carrier for immobilization of Cso2, a key enzyme 1 mL scale after ten reaction cycles, but the whole cells produced
for vanillin production, which is relatively labile, was first exam- only 2.5 mg (Fig. 2).
ined. The oxygenase Cso2 exhibits high activity in the pH range Since we confirmed that Cso2 immobilized on Sepabeads EC-EA
9.0–10.5, whereas the isoelectric point of Cso2 is pH 5.3 (Furuya had superior operational stability, we used the immobilized Cso2
et al., 2014). This property indicates that Cso2, with a negative catalyst in the cascade synthesis of vanillin from ferulic acid via
charge at the optimal pH, would be immobilized on anion-exchange 4-vinylguaiacol (Fig. 1B). We found that the first-step Fdc was as
carriers. Four anion-exchange carriers were tested. The matrix, easily immobilized on Sepabeads EC-EA as Cso2 was. We confirmed
functional group and particle size of each carrier are shown in that 5 mg of protein immobilized on 50 mg of the carrier efficiently
Table 1. When each carrier was incubated with Cso2 solution, converted ferulic acid to 4-vinylguaiacol (Fig. S3 in Supplementary
all were able to adsorb protein (0.04–0.10 mg mg-carrier−1 ) (Fig. data). Thus, the immobilized Fdc catalyst and the immobilized Cso2
S1A in Supplementary data). In addition, we examined vanillin- catalyst were added to a solution containing the substrate ferulic
producing activity of these carriers after adsorbing Cso2. When acid. Although a portion of the substrate and the product bound
each was incubated with isoeugenol, Cso2 immobilized on Sepa- to the carrier as described above, the immobilized Fdc and Cso2
beads EC-EA efficiently converted this substrate to vanillin (Fig. S1B efficiently converted ferulic acid to vanillin via 4-vinylguaiacol.
in Supplementary data). In contrast, Cso2 on the other carriers lost The immobilized enzymes produced 2.8 mM vanillin during 24 h
catalytic activity. These results suggested that Cso2 was compati- biotransformation (Fig. S4 in Supplementary data). After the first
ble with the polymethacrylate matrix and the ethylamine group of reaction cycle, the immobilized enzymes were collected from the
Sepabeads EC-EA (Table 1). reaction mixture and added to a fresh reaction mixture (Fig. 3).
We then attempted to produce vanillin from isoeugenol using In the second reaction cycle, the immobilized enzymes produced
Cso2 immobilized on Sepabeads EC-EA (Fig. 1A). The activity of the 3.5 mM vanillin. In the third and fourth reaction cycles, the immo-
immobilized Cso2 catalyst was compared with that of a whole- bilized enzymes maintained their activity, producing 3.1 mM and
cell catalyst, E. coli expressing Cso2. Each catalyst was incubated 2.4 mM vanillin, respectively. In subsequent reaction cycles, the
with a solution containing the substrate isoeugenol. The whole production of vanillin gradually decreased. We confirmed that the
cells produced 8.4 mM vanillin from 10 mM isoeugenol during 24 h intermediate 4-vinylguaiacol accumulated in the reaction mixture.
biotransformation (Fig. S2A in Supplementary data), whereas the The total amount of vanillin produced from ferulic acid by the
immobilized enzyme produced only 4.6 mM vanillin (Fig. S2B in immobilized Fdc and Cso2 catalysts reached 2.5 mg at the 1 mL scale
Supplementary data). We confirmed that a portion of the substrate after ten reaction cycles (Fig. 3). The amount of vanillin produced
isoeugenol and the product vanillin bound to the carrier, which from ferulic acid was lower than that produced from isoeugenol.
T. Furuya et al. / Journal of Biotechnology 243 (2017) 25–28 27

Table 1
Anion-exchange carriers tested in this study.

Sepabeads EC-EA Sepabeads EC-Q1A Diaion HPA25L Diaion WA21J

H2 CH3
C C H2 H H2 H
C O C C C C H2 H
O C C
CH2 CH3 CH3
CH2 CH2N+CH3Cl- CH2N+CH3Cl-

Structure NH2 CH3 CH3 CH2NH(CH2CH2NH)nH

Matrix polymethacrylate styrene-divinylbenzene styrene-divinylbenzene styrene-divinylbenzene


Functional group ethylamine quaternary alkylamine quaternary alkylamine polyamine
Particle size (␮m) 100–200 200–600 300–1200 300–1200

5 3 Author contributions

4 T.F., M.K. and K.K. designed research; T.F. and M.K. performed
research; M.K. analyzed data; and T.F., M.K. and K.K. wrote the

of vanillin (mg)
2
Vanillin (mM)

Total amount
3 paper.

2 Funding
1

1 This work was supported by a Novozymes Japan Research Fund


2016 to T.F.
0 0
1 2 3 4 5 6 7 8 9 10 Acknowledgements
Cycle number

We kindly thank H. Kawabata (Mitsubishi Chemical Group Sci-


Fig. 3. The vanillin production from ferulic acid by repetitive use of immobilized Fdc
and Cso2. Ferulic acid (10 mM) was incubated with the immobilized Fdc (5 mg pro- ence and Technology Research Center, Inc.) for the gift of samples
tein on 50 mg carrier) and the immobilized Cso2 (7 mg protein on 100 mg carrier) in of anion-exchange carries.
the reaction mixture (1 mL) for 24 h in each cycle. Vanillin produced in each cycle by
the immobilized enzymes (white bars) is shown. Total amount of vanillin produced
by the immobilized enzyme (white circles) is also shown. Data are the average of Appendix A. Supplementary data
three independent experiments, and error bars indicate standard deviation of the
mean. Supplementary data associated with this article can be found, in
the online version, at http://dx.doi.org/10.1016/j.jbiotec.2016.12.
021.

It is possible that the reaction time was not enough for the com- References
pletion of the two-step reaction in each cycle (Figs. S2B and S4 in
Supplementary data), or that the substrate ferulic acid and/or the Achterholt, S., Priefert, H., Steinbüchel, A., 2000. Identification of Amycolatopsis sp.
strain HR167 genes: involved in the bioconversion of ferulic acid to vanillin.
intermediate 4-vinylguaiacol had unfavorable effect on Cso2. Appl. Microbiol. Biotechnol. 54, 799–807.
In conclusion, we demonstrated that a coenzyme-independent Barghini, P., Di Gioia, D., Fava, F., Ruzzi, M., 2007. Vanillin production using
cascade consisting of the decarboxylase Fdc and the oxygenase Cso2 metabolically engineered Escherichia coli under non-growing conditions.
Microb. Cell Fact. 6, 13.
was easily applicable to an immobilized enzyme process for vanillin Barthelmebs, L., Diviès, C., Cavin, J.F., 2001. Expression in Escherichia coli of native
production. Although a lot of studies have been devoted to biotech- and chimeric phenolic acid decarboxylases with modified enzymatic activities
nological production of vanillin, this study is, to our knowledge, the and method for screening recombinant E. coli strains expressing these
enzymes. Appl. Environ. Microbiol. 67, 1063–1069.
first for vanillin production using immobilized enzymes. The immo-
Brady, D., Jordaan, J., 2009. Advances in enzyme immobilisation. Biotechnol. Lett.
bilized enzyme process provides new possibilities for the practical 31, 1639–1650.
bioproduction of vanillin, because vanillin can be prepared by only Di Gioia, D., Sciubba, L., Setti, L., Luziatelli, F., Ruzzi, M., Zanichelli, D., Fava, F., 2007.
mixing isoeugenol or ferulic acid with the immobilized enzyme(s) Production of biovanillin from wheat bran. Enzym. Microb. Technol. 41,
498–505.
that can be used repetitively. In the present study, as the activity Furuya, T., Miura, M., Kino, K., 2014. A coenzyme-independent
of Cso2 was not so high, a large amount of the enzyme was added decarboxylase/oxygenase cascade for the efficient synthesis of vanillin.
to the reaction mixture for the conversion of isoeugenol to vanillin. ChemBioChem 15, 2248–2254.
Furuya, T., Miura, M., Kuroiwa, M., Kino, K., 2015. High-yield production of vanillin
Accumulation of the intermediate 4-vinylguaiacol after the repeti- from ferulic acid by a coenzyme-independent decarboxylase/oxygenase
tive reaction with ferulic acid indicates that the second-step Cso2 two-stage process. New Biotechnol. 32, 335–339.
reaction limited further increase production of vanillin. We are Gallage, N.J., Møller, B.L., 2015. Vanillin-bioconversion and bioengineering of the
most popular plant flavor and its de novo biosynthesis in the vanilla orchid.
now investigating scale-up of the process as well as genetic engi- Mol. Plant 8, 40–57.
neering of Cso2 for higher activity and stability toward industrial Hilterhaus, L., Minow, B., Müller, J., Berheide, M., Quitmann, H., Katzer, M., Thum,
application. O., Antranikian, G., Zeng, A.P., Liese, A., 2008. Practical application of different
enzymes immobilized on sepabeads. Bioprocess Biosyst. Eng. 31, 163–171.
Hua, D., Ma, C., Song, L., Lin, S., Zhang, Z., Deng, Z., Xu, P., 2007. Enhanced vanillin
production from ferulic acid using adsorbent resin. Appl. Microbiol. Biotechnol.
74, 783–790.
Competing interests Kaur, B., Chakraborty, D., 2013. Biotechnological and molecular approaches for
vanillin production: a review. Appl. Biochem. Biotechnol. 169, 1353–1372.
Kloer, D.P., Schulz, G.E., 2006. Structural and biological aspects of carotenoid
All authors declare no competing or financial interests. cleavage. Cell Mol. Life Sci. 63, 2291–2303.
28 T. Furuya et al. / Journal of Biotechnology 243 (2017) 25–28

Lee, E.G., Yoon, S.H., Das, A., Lee, S.H., Li, C., Kim, J.Y., Choi, M.S., Oh, D.K., Kim, S.W., Sheldon, R.A., van Pelt, S., 2013. Enzyme immobilisation in biocatalysis: why, what
2009. Directing vanillin production from ferulic acid by increased acetyl-CoA and how. Chem. Soc. Rev. 42, 6223–6235.
consumption in recombinant Escherichia coli. Biotechnol. Bioeng. 102, 200–208. Walton, N.J., Mayer, M.J., Narbad, A., 2003. Molecules of interest: vanillin.
Ma, X.K., Daugulis, A.J., 2014. Effect of bioconversion conditions on vanillin Phytochemistry 63, 505–515.
production by Amycolatopsis sp. ATCC 39116 through an analysis of competing Wangrangsimagul, N., Klinsakul, K., Vangnai, A.S., Wongkongkatep, J., Inprakhon,
by-product formation. Bioprocess Biosyst. Eng. 37, 891–899. P., Honda, K., Ohtake, H., Kato, J., Pongtharangkul, T., 2012. Bioproduction of
Muschiol, J., Peters, C., Oberleitner, N., Mihovilovic, M.D., Bornscheuer, U.T., vanillin using an organic solvent-tolerant Brevibacillus agri 13. Appl. Microbiol.
Rudroff, F., 2015. Cascade catalysis–strategies and challenges en route to Biotechnol. 93, 555–563.
preparative synthetic biology. Chem. Commun. (Camb) 51, 5798–5811. Yamada, M., Okada, Y., Yoshida, T., Nagasawa, T., 2007. Biotransformation of
Narbad, A., Gasson, M.J., 1998. Metabolism of ferulic acid via vanillin using a novel isoeugenol to vanillin by Pseudomonas putida IE27 cells. Appl. Microbiol.
CoA-dependent pathway in a newly-isolated strain of Pseudomonas fluorescens. Biotechnol. 73, 1025–1030.
Microbiology 144, 1397–1405. Yang, J., Wang, S., Lorrain, M.J., Rho, D., Abokitse, K., Lau, P.C., 2009. Bioproduction
Priefert, H., Rabenhorst, J., Steinbüchel, A., 2001. Biotechnological production of of lauryl lactone and 4-vinyl guaiacol as value-added chemicals in two-phase
vanillin. Appl. Microbiol. Biotechnol. 56, 296–314. biotransformation systems. Appl. Microbiol. Biotechnol. 84, 867–876.
Rao, S.R., Ravishankar, G.A., 2000. Vanilla flavour: production by conventional and
biotechnological routes. J. Sci. Food Agric. 80, 289–304.
Rosazza, J.P., Huang, Z., Dostal, L., Volm, T., Rousseau, B., 1995. Review: biocatalytic
transformations of ferulic acid: an abundant aromatic natural product. J. Ind.
Microbiol. 15, 457–471.

You might also like