You are on page 1of 5

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/24493990

Biocatalytic Synthesis of Vanillin

Article in Applied and Environmental Microbiology · March 2000


DOI: 10.1128/AEM.66.2.684-687.2000 · Source: PubMed Central

CITATIONS READS
83 290

2 authors:

Tao Li John P N Rosazza


United States Environmental Protection Agency University of Iowa
20 PUBLICATIONS 769 CITATIONS 234 PUBLICATIONS 5,685 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Discovery/Biocatalysis with Nocardia iowensis View project

Development of Enzymatic Biosensor for Detecting Arsenic in Groundwater View project

All content following this page was uploaded by John P N Rosazza on 22 May 2014.

The user has requested enhancement of the downloaded file.


APPLIED AND ENVIRONMENTAL MICROBIOLOGY, Feb. 2000, p. 684–687 Vol. 66, No. 2
0099-2240/00/$04.00⫹0
Copyright © 2000, American Society for Microbiology. All Rights Reserved.

Biocatalytic Synthesis of Vanillin


TAO LI AND JOHN P. N. ROSAZZA*
Division of Medicinal and Natural Products Chemistry, and Center for Biocatalysis and
Bioprocessing, College of Pharmacy, University of Iowa, Iowa City, Iowa 52242
Received 1 July 1999/Accepted 17 November 1999

The conversions of vanillic acid and O-benzylvanillic acid to vanillin were examined by using whole cells and
enzyme preparations of Nocardia sp. strain NRRL 5646. With growing cultures, vanillic acid was decarboxylat-
ed (69% yield) to guaiacol and reduced (11% yield) to vanillyl alcohol. In resting Nocardia cells in buffer, 4-O-
benzylvanillic acid was converted to the corresponding alcohol product without decarboxylation. Purified No-
cardia carboxylic acid reductase, an ATP and NADPH-dependent enzyme, quantitatively reduced vanillic acid to
vanillin. Structures of metabolites were established by 1H nuclear magnetic resonance and mass spectral analyses.

Vanilla is one of the most widely used flavors in food indus- In this report, we describe biotransformations of vanillic acid
try. Natural vanilla extracted from the cured pods of the flow- by growing cells of Nocardia sp. strain NRRL 5646 and a pure
ers of Vanilla planifolia has an estimated net value of more aryl aldehyde oxidoreductase (carboxylic acid reductase) from
than $ 1 billion annually (18). Vanillin (3-methoxy-4-hydroxy- this organism. We also show that a complicating vanillic acid
benzaldehyde) is the most important organoleptic component decarboxylation reaction can be overcome by use of a vanillic
in vanilla. More than 12,000 tons of synthetic vanillin are pro- acid derivative as a substrate in whole-cell bioconversions.
duced each year from petrochemical and wood pulping indus-
tries (7). Strong market demand for natural and environmen- MATERIALS AND METHODS
tally friendly products has spawned efforts to produce vanillin
Biocatalysts. Nocardia species strain NRRL 5646 is maintained in the Uni-
by microbial transformation from natural substrates, including versity of Iowa College of Pharmacy culture collection and is grown and main-
phenolic stibenes (27), eugenol (19, 26), and ferulic acid (14, tained on slants of Sabouraud-dextrose agar or sporulating agar (ATCC no. 5
17). medium). The carboxylic acid reductase used in enzymatic reductions of vanillic
Ferulic acid [3-(4-hydroxy-3-methoxyphenyl)-propenoic acid was isolated from Nocardia sp. strain NRRL 5646 by our procedure (16).
The enzyme used in this work was pure, based on sodium dodecyl sulfate-
acid] is an extremely abundant plant product available from polyacrylamide gel electrophoresis analysis (15).
corn kernel hulls obtained from wet milling (21). Vanillic acid Chemicals. Vanillic acid (4-hydroxy-3-methoxybenzoic acid) (compound 1a
is the major product obtained by ␤-oxidation of ferulic acid by [Fig. 1]), vanillin (4-hydroxy-3-methoxybenzaldehyde) (compound 3a), vanillyl
species of Bacillus (6), Pseudomonas (11, 23), Polyporus (9), alcohol (4-hydroxy-3-methoxybenzyl alcohol) (compound 4a), guaiacol (2-me-
thoxyphenol) (compound 2), and benzyl bromide were purchased from Aldrich
Rhodotorula (8), and Streptomyces (17, 22). Thus, vanillic acid Chemical Co. (Milwaukee, Wis.). Other chemicals were purchased from Sigma
is an abundant, readily available precursor for the biocatalytic Chemical Co. (St. Louis, Mo.) and Aldrich unless otherwise indicated.
synthesis of vanillin. 4-O-Benzylation of vanillic acid was carried out by adding 11 ml of 1 N NaOH
Microbial reductions of carboxylic acids are widely observed to 900 mg of vanillic acid (5.4 mmol) in 16 ml of ethanol, followed by dropwise
addition of 1 g (5.8 mmol) of benzyl bromide in 2.5 ml of ethanol over 60 min.
in bacteria and fungi (1, 2, 4, 5, 10, 12, 13, 24, 25). One com- The reaction mixture was refluxed for 2 h before being poured into 150 ml of
mon mechanism of enzymatic carboxylic acid reduction in- water and acidified to pH 2 with 6 N HCl. The precipitate was collected by
volves activation of carboxyl groups with ATP to highly reac- filtration and crystallized in xylene to afford 802 mg of compound 1b (3.1 mmol,
tive carbonyl-AMP intermediates which are readily reduced by 56% yield). The structure of compound 1b was confirmed by low-resolution
electron impact mass spectrometry (EIMS) with the following parameters: m/z
NADPH to aldehydes (5, 13, 16) (see Fig. 1). Both the adeny- 258 (8%, M⫹) and 91 (100%, benzonium ion); 1H nuclear magnetic resonance
lation and carbonyl-AMP reduction steps are catalyzed by a (NMR) at 360 MHz (d6-acetone), ␦H 3.83 (3H, s, OCH3), 5.21 (2H, s, C-7⬘), 7.12,
single enzyme (5, 13, 16). A separate alcohol oxidoreductase 7.14 (1H, d, J ⫽ 8.3 Hz, H-5), 7.33 to 7.52 (5H, m, H-2⬘, -3⬘, -4⬘, -5⬘, and -6⬘), 7.58,
reduces the resulting aldehydes to their corresponding alcohol 7.57 (1H, d, J ⫽ 2.0 Hz, H-2), and 7.63 to 7.65 (1H, dd, J ⫽ 8.3, 2.0 Hz, H-6).
4-O-Benzylvanillin (compound 3b) was prepared by refluxing 32 g of vanillin
products (16). Although carboxylic acid reductases have been (210 mmol) with 26 ml (213 mmol) of benzyl bromide, and 30 g of potassium
purified from two species of Nocardia (13, 16), properties of bicarbonate in 200 ml of ethanol for 5 h. After cooling, the filtrate was crystal-
the reported enzymes are significantly different. lized at 4°C overnight. Recrystallization in ethanol gave 45.3 g (187 mmol, 85%
It is possible to envision an enzymatic synthesis of vanillin yield) of compound 3b confirmed by EIMS: m/z 242 (4%, M⫹), 91 (100%,
benzonium ion); 1H NMR at 360 MHz (CDCl3), ␦H 3.95 (3H, s, OCH3), 5.25
beginning either with ferulic acid (8, 17, 21) or with vanillic (2H, s, C-7⬘), 7.00, 6.98 (1H, d, J ⫽ 8.3 Hz, H-5), 7.34 to 7.43 (7H, m, H-2, -6, -2⬘,
acid as the starting materials. Such approaches are attractive -3⬘, -4⬘, -5⬘, and -6⬘), and 9.84 (1H, s, H-7).
because the natural carboxylic acid starting materials are abun- 4-O-Benzylvanillyl alcohol (compound 4b) was prepared by refluxing a mixture
dant and inexpensive and are soluble in aqueous media. Fur- of 15 g of potassium bicarbonate, 16 g (104 mmol) of vanillyl alcohol, and 18.3 g
of benzyl bromide (107 mmol) in 115 ml of ethanol for 3 h. The reaction workup
thermore, reductions of carboxylic acids to aldehydes are dif- was like that for 4-O-benzylvanillic acid to give pure product. Recrystallization
ficult to achieve by chemical means, and the reduction of from cyclohexane gave 17 g of compound 4b (64% yield). The low-resolution,
vanillic acid to vanillin has not been widely reported. fast atom bombardment (FAB) mass spectrum gave m/z 244 (17%, M⫹), 227
(9%, M⫹ ⫺ OH), and 91 (100%, benzonium ion); 1H NMR at 360 MHz
(CDCl3), ␦H 3.89 (3H, s, OCH3), 4.58 (2H, s, C-7), 5.16 (2H, s, H-7⬘), 6.76 to 6.79
(1H, dd, J ⫽ 8.3, 1.9 Hz, H-6), 6.82, 6.84 (1H, d, J ⫽ 8.2 Hz, H-5), 6.91, 6.92 (1H,
d, J ⫽ 1.7 Hz, H-2), and 7.26 to 7.43 (5H, m, H-2⬘, -3⬘, -4⬘, -5⬘, and -6⬘).
* Corresponding author. Mailing address: Division of Medicinal and
Chromatography. Thin-layer chromatography (TLC) was carried out on silica
Natural Products and Center for Biocatalysis and Bioprocessing, Col- gel GF254 plates (E. Merck, Darmstadt, Germany). Developed chromatograms
lege of Pharmacy, University of Iowa, Iowa City, IA 52242. Phone: were directly visualized under 254-nm UV light to observe fluorescence quench-
(319) 335-8842. Fax: (319) 335-8766. E-mail: john-rosazza@uiowa ing. Phenolic compounds were also visualized with Pauly’s reagent, which con-
.edu. sisted of solution A (0.5% sulfanilic acid in 2 N HCl), solution B (0.5% sodium

684
VOL. 66, 2000 BIOCATALYTIC SYNTHESIS OF VANILLIN 685

FIG. 1. Whole-cell and enzyme transformation pathways for vanillic acid and O-benzyl vanillic acid by Nocardia sp. strain NRRL 5646.

nitrite in water), and solution C (1 N potassium hydroxide in 50% ethanol- in distilled water and was adjusted to pH 7.2 with 6 N HCl before being auto-
water). Developed plates were first sprayed with a freshly prepared equal-volume claved at 121°C for 20 min. Cultures were incubated by shaking at 250 rpm at
mixture of solutions A and B and then sprayed with solution C before being 28°C on G25 Gyrotory shakers (New Brunswick Scientific Co., Edison, N.J.). A
heated with a heat gun to develop colors from yellow to orange. Aldehydes such 10% inoculum derived from a 72-h-old first-stage culture was used to initiate
as vanillin were also detected with 2,4,-dinitrophenylhydrazine (0.4% in 2 N HCl second-stage cultures, which were incubated as described above. After 24 h of
[wt/vol]) spray. With dichloromethane-acetonitrile-formic acid (75:25:1 [vol/vol/ incubation in the second stage, 160 mg of vanillic acid (in 2 ml of dimethyl
vol]), Rf values for vanillic acid, vanillin, vanillyl alcohol, and guaiacol were 0.5, sulfoxide) was added to each of three flasks. Control cultures included everything
0.4, 0.8, and 0.9, respectively. With dichloromethane-acetic acid-benzene (100: except vanillic acid.
2:2 [vol/vol/vol]), Rf values for 4-O-benzylvanillic acid (compound 1b), 4-O- Samples were removed at various time intervals, adjusted to pH 2 with 6 N
benzylvanillin (compound 3b), and 4-O-benzylvanillyl alcohol (compound 4b) HCl, extracted with 1 ml of ethyl acetate, and centrifuged for 1 min at 2,500 rpm.
were 0.7, 0.8, and 0.2, respectively. Organic layers were removed, evaporated to dryness, and reconstituted in 0.5 ml
High-performance liquid chromatography (HPLC) was performed with a Shi- of methanol. The samples were spotted onto TLC plates for analysis.
madzu liquid chromatograph equipped either with a LC-10AD system controller, Samples of 3 ml were also taken over the course of vanillic acid transformation
four FCV-10AL pumps, and a photodiode array UV-Vis detector (SPD-M6A) or for HPLC analysis. To each sample, 200 ␮l of o-anisic acid solution (30 mg/ml in
with a SCL 6-B system controller, two LC-6A pumps, and a variable-wavelength acetonitrile) was added as an internal standard. Culture samples were then
UV detector. Eluted peaks were detected at 273 to 276 nm or at 290 nm and acidified to pH 2 with 6 N HCl, 1 ml was loaded onto solid-phase extraction
identified by comparison with authentic compounds. Separations carried out cartridges (Chem Elut CE 1001, 1-ml aqueous solution capacity; Varian, Harbor
under isocratic conditions over a Versapack C18 column (250 by 4.6 mm, 10-␮m City, Calif.). After 5 min, cartridges were extracted twice with 3 ml each of
particle size; Alltech, Deerfield, Ill.) with acetonitrile-water-formic acid (20:80:1) dichloromethane-acetonitrile (90:10). These organic extracts were combined and
at a flow rate of 0.9 ml/min gave retention volumes (Rv) of 5.0 ml for vanillyl used for HPLC analysis.
alcohol, 8.1 ml for vanillic acid, 13.2 ml for vanillin, 16.1 ml for o-anisic acid, and The microbial reaction was terminated 40 h after the addition of vanillic acid.
18.7 ml for guaiacol. Standard curves for vanillic acid (compound 1a), vanillin The culture was acidified to pH 2 with 6 N HCl. After centrifugation to remove
(compound 3a), vanillyl alcohol (compound 4a), and guaiacol (compound 2) cells and other solids, the supernatant from 200 ml of culture was loaded onto a
were established over the range of 0.3 to 8 ␮g. For quantitation of 4-O-benzyl- solid-phase extraction cartridge (Chem Elut 1200, 200-ml aqueous solution ca-
vanillic acid biotransformations, separations were carried out under isocratic pacity). After 5 min of equilibration, the cartridge was washed three times with
conditions over a 10 ␮m C18 Versapack column (300 by 4.1 mm; Alltech) with 200 ml of hexane and then three times with 200 ml of dichloromethane. The
acetonitrile-water-formic acid (30:70:1) at a flow rate of 2 ml/min. Rv values were organic extracts were separately concentrated by rotary evaporation to give 45
8.1 ml for p-anisic acid, 21.9 ml for 4-O-benzylvanillyl alcohol (compound 4b), mg of oil (from hexane) and 90 mg of residue (from dichloromethane). The
33.4 ml for 4-O-benzylvanillic acid (compound 1b), and 61.1 ml for 4-O-benzyl- dichloromethane extracts were streaked onto a 1-mm-thick preparative TLC
vanillin (compound 3b). Standard curves for these compounds were established plate (20 by 20 cm), which was developed in dichloromethane-acetonitrile-formic
over the range of 0.3 to 3 ␮g. acid (75:25:1 [vol/vol]). The separated bands after development were scraped
Spectroscopy. Mass spectra were obtained by using a Trio-1 MS linked with a from the plate, and compounds were eluted from silica gel with a mixture of
5890 Hewlett-Packard gas chromatograph. EIMS was performed at an ionization dichloromethane and acetonitrile (70:30 [vol/vol]). The band extracts were
voltage of 70 eV. For direct inlet probe analysis, the probe temperature was set checked for purity by TLC, and concentrated for mass spectrometry and 1H
at 30°C for 1 min, raised to 300°C at 150°C/min, and held at 300°C for 10 min for NMR spectroscopy.
analysis. For gas chromatography-mass spectrometry, separations were carried Reduction of vanillic acid to vanillin by Nocardia carboxylic acid reductase.
out on an OV-1 capillary column (10 m by 0.25 mm; 1-␮m film thickness) with The reduction was carried out in a reaction mixture of 200 ml of 50 mM Tris-HCl
helium as carrier gas at a flow rate of 20 ml/min. The column oven temperature buffer (pH 7.5) containing 34 mg of vanillic acid, 59 mg of NADPH, 110 mg of
was held at 50°C for 1 min, raised to 250°C at a rate of 15°C/min, and held at that ATP, and 100 ␮g of purified carboxylic acid reductase (0.6 U) (16). The reaction
temperature for 10 min. Injector and detector temperatures were 220 and 270°C, mixture was incubated at 30°C with gentle shaking (50 rpm) for 24 h after which
respectively. FAB mass analyses were performed by using a ZAB-HF mass the entire reaction mixture was loaded onto a solid-phase extraction cartridge
spectrometer (VG Analytical, Inc.). Ionizing matrices were either 3-nitrobenzyl (Chem Elut 1200, 200-ml aqueous solution capacity). After 5 min of equilibra-
alcohol or thioglycerol. tion, the cartridge was washed twice with 100 ml of hexane, followed by two
NMR spectra were obtained with a Bruker WM 360-MHz high-field spectrom- washes 100 ml of dichloromethane. The dichloromethane solution was concen-
eter equipped with an IBM Aspect-2000 processor. Tetramethylsilane was used trated to less than 2 ml. The solution was transferred to a vial, and the solvent
as the internal standard. Chemical-shift values are reported in parts per million was removed with a stream of nitrogen to give pure vanillin as determined by
(ppm), and coupling constants (J values) are given in hertz. Abbreviations for TLC, EIMS, and 1H NMR analyses.
NMR are as follows: s, singlet; d, doublet; t, triplet; dd, doublet of doublets; and Nocardia resting cell conversions of 4-O-benzylvanillic acid. Nocardia sp. was
m, multiplet. grown by using the standard two-stage incubation protocol. After 24 h in the
Biotransformation of vanillic acid by Nocardia isolates. A preparative scale second-stage culture, 5 mg of benzoic acid per ml was added as an inducer for
incubation was grown by our standard two-stage incubation protocol (3) in carboxylic acid reductase (16). The cells were harvested by centrifugation at
200-ml volumes of sterile soybean flour-glucose medium held in stainless-steel- 8,000 ⫻ g for 20 min at 24 h after the benzoate addition. Cells were washed twice
capped, 1-liter, DeLong culture flasks. The medium contained 20 g of glucose, with 0.9% saline solution and pelleted before use.
5 g of yeast extract, 5 g of soybean flour, 5 g of NaCl, and 5 g of K2HPO4 per liter For preparative scale biotransformations, 2.8 g of cells (wet weight) were
686 LI AND ROSAZZA APPL. ENVIRON. MICROBIOL.

M⫹ ⫺ CH3 ⫺ H2O), and 81 (100%, M⫹ ⫺ CH3 ⫺ CO), and


an 1H NMR spectrum at 360 MHz (CDCl3), ␦H 3.76 (3H, s,
OCH3), 6.80, 6.85 (3H, m, H-3, H-4, H-6), and 6.90 to 6.93
(1H, m, H-5). These properties were identical to those for an
authentic sample of guaiacol.
The dichloromethane extract gave two compounds (4 mg
each) by preparative TLC. The compound at Rf 0.77 was spec-
trally identical to authentic vanillin (compound 3a): low-reso-
lution EIMS, m/z 152 (88%, M⫹), 151 (100%, M⫹ ⫺ H), 137
(6%, M⫹ ⫺ CH3), 123 (13%, M⫹ ⫺ CHO); 1H NMR at 360
MHz (CDCl3), ␦H 3.96 (3H, s, OCH3), 6.22 (1H, s, OH), 7.03,
7.06 (1H, d, J ⫽ 8.5, H-5), 7.26 to 7.44 (2H, m, H-2, H-6), and
9.83 (1H, s, CHO). The compound at Rf ⫽ 0.35 was identical
to vanillyl alcohol (compound 4a): low-resolution EIMS, m/z
154 (28%, M⫹), 136 (50%, M⫹ ⫺ H2O); 1H NMR at 360 MHz
(CDCl3), ␦H 3.90 (3H, s, OCH3), 4.61 (2H, s, CH2OH), and
6.82 to 6.92 (3H, m, H-2, H-5, H-6).
Reduction of vanillic acid by purified Nocardia carboxylic
acid reductase. The enzymatic reaction gave 8 mg (53 ␮mol) of
vanillin by TLC, EIMS, and 1H NMR analyses, all of which
were identical to an authentic sample of vanillin. Based upon
the estimated amount of NADPH consumed (51 ␮mol,
FIG. 2. Time course of growing cell biotransformations of vanillic acid by
[change in optical density at 340 nm]) during the course of the
Nocardia sp. strain NRRL 5646. Triplicate sample averages for vanillic acid (F), reaction, the yield of vanillin was essentially quantitative. The
vanillyl alcohol (■), vanillin (䊐), and guaiacol (E) had variances of not more remainder of the vanillic acid substrate was unreacted.
than ⫾5%. Transformation of 4-O-benzylvanillic acid with resting cells
of Nocardia sp. By HPLC, benzoate-induced cells of Nocardia
suspended in 400 ml of 50 mM Tris HCl (pH 7.4) containing 1% glucose, 10 mM
sp. rapidly transformed 4-O-benzylvanillic acid to metabolites
MgCl2, and 5 mM phosphate. 4-O-Benzylvanillic acid (0.7 g in 2 ml of dimethyl within 40 min (Fig. 3). In the first 15 min, the aldehyde 3b
sulfoxide) was added to the cell suspension, which was incubated with shaking accumulated in transient fashion to a maximum 30% yield (175
at 150 rpm at 30°C for 15 min. The reaction mixture was acidified to pH 2, ␮g/ml) after which compound 3b was further quantitatively
centrifuged to remove cells and other unwanted solids, and loaded onto two
solid-phase extraction cartridges (Chem Elut 1200, 200 ml on each). After 5 min
reduced to the corresponding alcohol 4b.
of equilibration, each cartridge was washed three times with 200 ml of ethyl In the preparative scale incubation reaction, TLC analysis
acetate. The ethyl acetate solution was concentrated by rotary evaporation to indicated that benzoate-induced Nocardia cells produced two
give 670 mg of crude product, which was reconstituted to 2.0 ml with dichlo- major metabolites corresponding to 4-O-benzylvanillin (com-
romethane and loaded onto a silica gel column (0.9 by 20 cm). 4-O-Benzylvanillin
(170 mg, 24% yield) was eluted with dichloromethane, while 4-O-benzylvanillyl
pound 3b, 170 mg, 24% yield) and 4-O-benzylvanillyl alcohol
alcohol (210 mg, 30% yield) was eluted with dichloromethane-benzene-acetic (compound 4b, 210 mg, 30% yield). These products were spec-
acid (100:4:6 [vol/vol/vol]). These compounds were spectrally (EIMS and NMR)
and chromatographically identical to synthetic standards described earlier.
4-O-Benzylvanillin obtained from the preparative-scale incubation was dis-
solved in 4 ml of ethanol, 5 ml of concentrated HCl was added, and the mixture
was refluxed for 2 h. The solvent was removed by rotary evaporation, and the
residue dissolved in 0.5 ml was purified by silica gel column chromatography (0.9
by 5 cm) eluted with dichloromethane to give 65 mg of pure vanillin as deter-
mined by TLC, EIMS, and 1H NMR.
Analytical scale biotransformations were conducted in the same medium as for
preparative biotransformations. Suspensions of cells (0.05 g [wet weight]/ml)
containing 0.6 mg of 4-O-benzylvanillic acid per ml were incubated at 30°C.
Samples (1 ml) taken during the course of incubation received 50 ␮l of p-anisic
acid (4 mg/ml in methanol) as an internal standard. Samples were acidified to pH
2 with 6 N HCl and loaded onto solid-phase extraction cartridges (Chem Elut CE
1001), which were equilibrated for 5 min and then eluted twice with 3 ml of ethyl
acetate. Pooled eluates were analyzed by HPLC.

RESULTS
Biotransformation of vanillic acid with growing cells of No-
cardia sp. HPLC analyses of vanillic acid biotransformation by
Nocardia sp. (Fig. 2) showed that almost all vanillic acid was
consumed within 48 h to give guaiacol (69% yield) by decar-
boxylation and vanillyl alcohol (11% yield) by carboxylic acid
reduction. Only traces of vanillin were detected in this analyt-
ical experiment. To properly characterize vanillic acid metab-
olites, a preparative scale incubation gave 45 mg of a sample
from the hexane extract of a 40-h biotransformation culture.
The product from the hexane extract gave a single spot by TLC
FIG. 3. Time course of resting cell biotransformations of O-benzylvanillic
(Rf ⫽ 0.89), was orange with Pauly’s phenolic spray reagent, acid by Nocardia sp. strain NRRL 5646. Triplicate sample averages for O-ben-
gave an EIMS spectrum data as follows: m/z (percent relative zylvanillic acid (F), O-benzylvanillyl alcohol (■), and O-benzylvanillin (䊐) had
abundance) 124 (78%, M⫹), 109 (92%, M⫹ ⫺ CH3), 91 (44%, variances of not more than ⫾5%.
VOL. 66, 2000 BIOCATALYTIC SYNTHESIS OF VANILLIN 687

trally and chromatographically comparable to synthetic com- (unpublished data). We are now exploring the possibility of clon-
pounds 3b and 4b. Thus, 4-O-benzylvanillic acid was reduced ing the carboxylic acid reductase in a suitable biocatalytic host.
first to the aldehyde 3b and subsequently to the alcohol 4b
(Fig. 1). To confirm the identity of compound 3b, chemical
ACKNOWLEDGMENTS
removal of the benzyl group from the microbial metabolite
gave 65 mg of vanillin identified by spectral and chromato- Tao Li acknowledges support through a Center for Biocatalysis and
graphic comparisons with authentic vanillin. Bioprocessing Fellowship, and we are grateful for financial support
from USDA through the Byproducts for Biotechnology Consortium.
DISCUSSION
REFERENCES
Many microbial transformation approaches have been used 1. Arfman, H. A., and W. R. Abraham. 1993. Microbial reduction of aromatic
to produce vanillin. In most cases, the yields of vanillin were carboxylic acids. Z. Naturforsch. 48c:52–57.
2. Bachman, D. M., B. Dragoon, and S. John. 1960. Reduction of salicylate to
low, and times for biotransformation reactions are long (7, 14, saligenin by Neurospora. Arch. Biochem. Biophys. 91:326.
17, 19, 26, 28). 3. Betts, R. E., D. E. Walters, and J. P. Rosazza. 1974. Microbial transforma-
The synthesis of vanillin from ferulic acid has been reported tions of antitumor compounds. 1. Conversion of acronycine to 9-hydroxya-
with several microorganisms (6, 9, 17, 22, 23). In P. acidivorans cronycine by Cunninghamella echinulata. J. Med. Chem. 17:599–602.
4. Chen, Y., and J. P. N. Rosazza. 1994. Microbial transformation of ibuprofen
(23), manometric studies suggested that vanillin was produced by a Nocardia species. Appl. Environ. Microbiol. 60:1292–1296.
by a two-step reaction including water addition to the cin- 5. Gross, G. G. 1972. Formation and reduction of intermediate acyl-adenylate
namoyl side chain and subsequent retro-aldol condensation to by aryl-aldehyde NADP oxidoreductase from Neurospora crassa. Eur. J. Bio-
give vanillin. In general, yields were low because vanillin was chem. 31:585–592.
6. Gurujeyalakshmi, G., and A. Mahadevan. 1987. Dissimilation of ferulic acid
further oxidized to vanillic acid which was O-demethylated to by Bacillus subtilis. Curr. Microbiol. 16:69–73.
protocatechuic acid or decarboxylated to guaiacol. The addi- 7. Hagedorn, S., and B. Kaphammer. 1994. Microbial biocatalysis in the gen-
tion of sulfhydryl compounds to ferulic acid transformation me- eration of flavor and fragrance chemicals. Annu. Rev. Microbiol. 48:773–800.
dium increased vanillin accumulations by Pseudomonas putida 8. Huang, Z., L. Dostal, and J. P. N. Rosazza. 1993. Mechanisms of ferulic acid
conversions to vanillic acid and guaiacol by Rhodotorula rubra. J. Biol. Chem.
ATCC 55180 (14). However, the mechanism by which sulfhy- 268:23954–23958.
dryl reagents improved vanillin yields was not clear. Mulheim 9. Ishikawa, H., W. J. Schubert, and F. F. Nord. 1963. The degradation by
and Lerch recently described a relatively high yield direct con- Polyporus versicolor and Formes fomentarius of aromatic compounds struc-
version of ferulic acid to vanillin by cultures of Streptomyces turally related to softwood lignin. Archiv. Biochem, Biophys. 100:140–149.
10. Jezo, I., and J. Zemek. 1986. Enzymatische Reducktion einiger aromatischer
setonii (17). With this organism, yields of vanillin were directly Carboxysäuren. Chem. Papers 40:279–281.
proportional to ferulic acid concentrations in culture media. 11. Jurková, M., and M. Wurst. 1993. Biodegradation of aromatic carboxylic
In our laboratory, the mechanism of ferulic acid transforma- acids by Pseudomonas mira. FEMS Microbiol. Lett. 111:245–250.
tion to vanillic acid by R. rubra IFO 889 occurred by ␤-oxida- 12. Kato, N., H. Konishi, K. Uda, M. Shimao, and C. Sakazawa. 1988. Microbial
reduction of benzoate to benzyl alcohol. Agric. Biol. Chem. 52:1885–1886.
tion (8). The ready availability of vanillic acid by this process, 13. Kato, N., E. H. Joung, H. C. Yang, M. Masuda, M. Shimao, and H. Yanase.
and the discovery of a vanillic acid reduction pathway in No- 1991. Purification and characterization of aromatic acid reductase from
cardia sp. strain NRRL 5646 suggested the potentials for whole Nocardia asteroides JCM 3016. Agric. Biol. Chem. 55:757–762.
cells or the pure carboxylic acid reductase from this microor- 14. Labuda, I. M., S. K. Goers, and K. A. Keon. July 1992. Bioconversion process
for the production of vanillin. U.S. patent 5,128,253.
ganism for vanillin synthesis. 15. Laemmli, U. K. 1970. Cleavage of structural proteins during the assembly of
With pure Nocardia carboxylic acid reductase, the ATP- and the head of bacteriophage T4. Nature (London) 227:680–685.
NADPH-dependent reduction of vanillic acid was quantitative, 16. Li, T., and J. P. N. Rosazza. 1997. The purification, characterization, and
yielding only vanillin and no complicating byproducts. In whole- properties of an aldehyde oxidoreductase from Nocardia sp. NRRL 5646. J.
Bacteriol. 179:3482–3487.
cell Nocardia biotransformation reactions, vanillic acid decar- 17. Mulheim, A., and K. Lerch. 1999. Towards a high-yield bioconversion of
boxylation to guaiacol was the major complicating pathway. ferulic acid to vanillin. Appl. Microbiol. Biotechnol. 51:456–461.
The identification of guaiacol and vanillyl alcohol as metabo- 18. Prince, R. C., and D. E. Gunson. 1994. Just plain vanilla? Trends Biol. Sci.
lites confirmed that Nocardia sp. strain NRRL 5646 possesses 19:521.
19. Rabenhorst, J., and R. Hopp. May 1991. Process for the preparation of
two different metabolic pathways for the biotransformation of vanillin. U.S. patent 5,017,388.
vanillic acid. These pathways are (i) decarboxylation to guaia- 20. Raman, T. S., and E. R. B. Shanmugasundaram. 1962. Metabolism of some
col and (ii) reduction to vanillin and further reduction to va- aromatic acids by Aspergillus niger. J. Bacteriol. 84:1340–1341.
nillyl alcohol (Fig. 1). 21. Rosazza, J. P. N., Z. Huang, L. Dostal, T. Volm, and B. Rousseau. 1995.
Biocatalytic transformation of ferulic acid: an abundant aromatic natural
We also observed vanillic acid decarboxylation in Rhodo- product. J. Ind. Microbiol. 15:457–471.
torula rubra (8) and in Bacillus pumilis (21). Deuterium isotope 22. Sutherland, J. B., D. L. Crawford, and A. L. Pometto III. 1983. Metabolism
incorporation studies revealed that the decarboxylation reac- of cinnamic, p-coumaric, and ferulic acids by Streptomyces, setonii. Can. J.
tion involved the enzymatic tautomerization of vanillic acid Microbiol. 29:1253–1257.
23. Toms, A., and J. M. Wood. 1970. The degradation of trans-ferulic acid by
through the para-hydroxyl group to a quinoid intermediate be- Pseudomonas acidovorans. Biochemistry 9:337–343.
fore decarboxylation. Blocking the p-phenolic group was con- 24. Tsuda, Y., K. Kawai, and S. Nakajima. 1984. Asymmetric reduction of 2-
sidered as a possible means of preventing the enzymatic tau- methyl-2-aryloxyacetic acids by Glomerella cingulata. Agric. Biol. Chem. 48:
tomerization reaction and subsequent decarboxylation. To test 1373–1374.
25. Tsuda, Y., K. Kawai, and S. Nakajima. 1985. Microbial reduction of 2-phe-
this hypothesis, we employed the benzylic functional group, nylpropionic acid, 2-benzyloxypropionic acid and 2-(2-furfuryl)propionic
which is readily removed by simple acid treatment. However, acid. Chem. Pharm. Bull. 33:4657–4661.
other alternative blocking groups, such as esters, could func- 26. Washisu, Y., A. Tetsushi, N. Hashimoto, T. Kanisawa. 1993. Manufacture of
tion equally well. 4-O-Benzylvanillic acid was not decarboxy- vanillin and related compounds with Pseudomonas. Japanese Patent No.
5,227,980
lated and was quantitatively reduced to 4-O-benzylvanillin and 27. Yoshimoto, T., M. Samejima, N. Hanyu, and T. Koma. August 1990. Dioxy-
4-O-benzylvanillyl alcohol. We know that Nocardia sp. also genase for styrene cleavage manufactured by Pseudomonas. Japanese patent
contains another reductase that converts aldehydes to alcohols 2,195,871.

View publication stats

You might also like