Anne 1999

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Can the Combination of Electrochemical

Regeneration of NAD+, Selectivity of


L-␣-Amino-Acid Dehydrogenase, and
Reductive Amination of ␣-Keto-Acid Be
Applied to the Inversion of Configuration
of a L-␣-Amino-Acid?

Agnès Anne,1 Christian Bourdillon,2 Sandra Daninos,1 Jacques Moiroux1


1
The Laboratoire d’Electrochimie Moléculaire, Unité Mixte de Recherche
Université–CNRS No. 7591, Université de Paris 7, Denis Diderot, 2 Place
Jussieu, 75251 Paris Cedex 05, France; telephone: +33 1 44 27 28 01; Fax:
+33 1 44 27 76 25; e-mail: moiroux@paris7.jussieu.fr
2
The Laboratoire de Technologie Enzymatique, UPRES-A No 6022,
Université de Technologie de Compiègne, BP 20529,
60205 Compiègne Cedex, France
Received 19 June 1998; accepted 5 December 1998

Abstract: The inversion of configuration of L-alanine can or by enzymatic systems (Cosnier et al., 1997; Palmore et
be carried out by combining its selective oxidation in the al., 1998). The direct electrochemical regeneration can be
presence of NAD+ and L-alanine dehydrogenase, electro-
chemical regeneration of the NAD+ at a carbon felt an- performed with a yield exceeding 99.99%, but requires an
ode, and reductive amination of pyruvate, i.e., reduction anodic potential that may be too positive for the selective
of its imino derivative at a mercury cathode, the reaction oxidation of some substrates (Bonnefoy et al., 1988). The
mixture being buffered with concentrated ammonium/ overpotential can be lowered markedly by the mediation.
ammonia (1.28M / 1.28M). The dehydrogenase exhibits
astonishing activity and stability under such extreme Thus, it is tempting to make continuous use of the electro-
conditions of pH and ionic strength. chemical regeneration for preparative processes (Somers et
The main drawback of the process is its slowness. At al., 1997).
best, the complete inversion of a 10 mM solution of L- The oxidation of secondary alcohols and amines by
alanine requires 140 h. A careful and detailed quantita-
tive analysis of each of the key steps involved shows that
NAD+ to ketones in the presence of the relevant dehydro-
the enzyme catalyzed oxidation is so thermodynamically genases is thermodynamically uphill (Lee and Whitesides,
uphill that it can be driven efficiently to completion only 1985). In practical reactors the sole driving force provided
when both the coenzyme regeneration and the pyruvate by the electrochemical regeneration of the coenzyme is not
reduction are very effective. The first condition is easily sufficient to draw the oxidation toward completion even
fulfilled. Under the best conditions, it is the rate of the
chemical reaction producing the imine which controls when the configuration of the system is optimized (Fassou-
the whole process kinetically. © 1999 John Wiley & Sons, ane et al., 1990; Lortie et al., 1992). The only effective
Inc. Biotechnol Bioeng 64: 101–107, 1999. method is to remove the product as it forms (Lee and
Keywords: electrochemical; regeneration; NAD; amino Whitesides, 1985). Several examples of alcohol dehydroge-
acid; configuration
nase catalyzed oxidations of alcohols by NAD+ to obtain
INTRODUCTION chiral products have been reported, various methods being
used for the regeneration of NAD+ (Hilt et al., 1997a,b;
It is now well established that the electrochemical regen-
Jones, 1985; Lee and Whitesides, 1985).
eration of enzymatically active NAD+ can be achieved quite
There exist only two possibilities of applying the regen-
efficiently either directly at a carbon anode (Bonnefoy et al.,
eration of NAD+ to the preparation of a D-␣-alcohol or
1988; Fassouane et al., 1990; Laval et al., 1987) or mediated
amino acid. In the first case, it may be envisaged to proceed
by polyoxometalates (Essaadi et al., 1994; Keita et al.,
to the selective oxidation of the L-enantiomer of a racemic
1995; Keita et al., 1996; Sadakane and Steckhan, 1998;
mixture in the presence of the L-enzyme and end up with a
Steckhan, 1994), quinonoid compounds (Blankespoor and
mixture of the carbonyl product and the D-enantiomer. In
Miller, 1984; Tse and Kuwana, 1978; Persson et al., 1985)
and their transition-metal complexes (Hilt et al., 1997a,b), the second case, the selective oxidation of the L-enantiomer
is coupled to a non-enzymatically catalyzed reduction of the
ketone which produces both enantiomers. As a result, half a
Correspondence to: Jacques Moiroux molecule of L-enantiomer is transformed into half a mol-

© 1999 John Wiley & Sons, Inc. CCC 0006-3592/99/010101-07


ecule of D-enantiomer each time a molecule of NAD+ is be reached practically. Such an optimized limit appears too
regenerated. Then one can start either with the L-enantiomer low for technical application.
or with the ketone and transform it completely into the sole
D-enantiomer. Previously, the inversion of configuration of
MATERIALS AND METHODS
L-lactate was thus achieved, in the presence of L-lactate
dehydrogenase, by coupling the electrochemical regenera-
tion of NAD+ at an anode to the electrochemical reduction Materials
of pyruvate at a cathode (Biade et al., 1992). In the present Enzymes and biochemicals, including L-alanine, D-alanine
work, we show that the same approach can be applied to the and D, L-alanine were purchased from Sigma, Saint Quen-
inversion of configuration of a L-␣-amino acid provided tin Follonier, France, and used without further purification.
that the medium is buffered with concentrated NH3/NH4+ L-alanine dehydrogenase (EC 1. 4. 1. 1) and D-amino-acid
and that the enzyme is stable and still satisfactorily active oxidase (EC 1. 4. 3. 3) and L-lactic acid dehydrogenase (EC
under such extreme conditions of pH and ionic strength. The 1. 1. 1. 27) were from Bacillus subtilis, porcine kidney, and
scheme of the process is given in Figure 1. rabbit muscle, respectively. All other chemicals and the Na-
L-alanine is the starting ␣-amino acid. L-alanine dehy- fion perfluorinated membranes separating the working elec-
drogenase (L-AlaDH) is the enzyme which catalyzes its trode and auxiliary electrode compartments inside each re-
oxidation into pyruvate. Like other ␣-keto carboxylates, py- actor were from Aldrich, Saint Quentin Follonier, France.
ruvate and its imine exist at equilibrium in the presence of The NH3/NH4+ (1.28M / 1.28M) buffer was obtained by
ammonia. We take advantage of the fact that the electro- dissolution of (NH4)2 SO4 and concentrated aqueous am-
chemical reduction of the imine is markedly easier and pro- monia (14.7M) from Prolabo in distilled water. The felt of
duces the racemic 0.5 L-alanine + 0.5 D-alanine. Therefore, carbon fibers (RVG 4000) was obtained from Carbone-
the cathode potential can be tuned so as to achieve a selec- Lorraine, Geneillies, France.
tive reduction (Anne et al., 1994; Jeffery and Meisters,
1978; Lund, 1970).
The key steps of the reaction scheme involved in Figure Apparatus
1 are the electrochemical regeneration of NAD+ which is For the study of enzyme kinetics, the NADH concentration
irreversible and rapid, provided that the anode potential is was determined by measuring the absorbance at 340 nm (␧
positive enough (Biade et al., 1992; Bonnefoy et al., 1988; ⳱ 6320 M−1 cm−1) with a Hewlett-Packard HP 8452 spec-
Fassouane et al., 1990; Laval et al., 1987), the reversible trophotometer. The potentials of the working electrodes vs.
enzymatically catalyzed reaction, the reversible formation SCE electrodes were controlled with home-made potentio-
of the imine, and the irreversible electrochemical reduction stats. The flow rates of the circulating solutions were con-
of the latter (Anne et al., 1994). The flow of two moles of trolled with a Gilson Minipuls (Villius le Bel, France) 3
electrons through the system provokes the transformation of peristaltic pump equipped with a 4-channel head.
half a mole of L-alanine into the same quantity of D-alanine
while one mole of NAD+ undergoes a complete redox cycle.
Besides demonstrating that the inversion of configuration of Reactors
the amino acid is feasible in such a process, the quantitative The reactor in which NAD+ was regenerated was as previ-
analysis of its kinetic behavior allows for the determination ously described (Fassouane et al., 1990). It was used in the
of the steps which control the overall rate of D-enantiomer vertical position, the solutions flowing from bottom to top.
production. As a result, the important parameters can be The reduction reactor was identical but used in the horizon-
identified, and it is possible to show how the rate could be tal position, the carbon felt being replaced by the mercury
improved and thus, evaluate the optimized limit that could pool. The reactors were fed as previously described (Biade
et al., 1992).

Enzymatic Assays
The method for the enzymatic assay of L-alanine was es-
sentially similar to that described by Williamson (1985).
Exhaustive catalyzed oxidation was achieved in the pres-
ence of NAD+ and hydrazine in excess. The determination
was carried out as follows: 0.3 mL of sample containing
L-alanine and 2.5 mL of 0.2M sodium pyrophosphate buffer
containing 0.48M hydrazine and 1.25 mM EDTA (pH 9.2)
were mixed in a 1 cm spectrophotometric cell. After a 10
min standing at 25°C, NAD+ (0.1 mL of a 60 mM solution
Figure 1. Scheme of the process. The starting materials can be either in H2O) was added, and the contents were mixed again. The
L-alanine, racemic D,L-alanine or pyruvate, NAD+ or NADH. enzymatic reaction was started by addition of the solution of

102 BIOTECHNOLOGY AND BIOENGINEERING, VOL. 64, NO. 1, JULY 5, 1999


L-alanine dehydrogenase (0.1 mL of a 0.9 mg/mL solution). and Whitesides,1981). We chose to operate in the NH3/
L-alanine was quantified from the increase of absorbance at NAD4+ (1.28M / 1.28M) buffer (pH ⳱ 9.2) the counter ion
340 nm after completion of the reaction (approximately 60 being sulfate. At a greater buffer concentration, the solution
min). The enzymatic determination of L-alanine was unaf- becomes too viscous.
fected by the presence of D-alanine in the sample.
The enzymatic assay of D-alanine was performed accord-
Rate of the Electrochemical Reduction of
ing to a previously described procedure (Gra␤l and Supp, the Imine
1985) that involves complete dioxygen D-amino acid oxi-
dase catalyzed oxidation to pyruvate followed by NADH The flow-through reactor is of the same type as the one
L-lactate dehydrogenase catalyzed reduction. Once cor- already used for the inversion of L-lactate (Biade et al.,
rected for the blank test, the decrease in absorbance at 340 1992). It has a volume vred of 10 mL. To test its efficiency,
nm was used for the quantitative determination of D-alanine a 2 mM solution of pyruvate in the NH3/NH4+ (1.28M /
which was unaffected by the presence of L-alanine. 1.28M) buffer (pH ⳱ 9.2) plus 0.5 mM NAD+ and 0.1
mg/mL L-AlaDH is pumped into the reactor at a flow rate
of 20 mL/h. At the output, D- and L-alanine (see Materials
RESULTS AND DISCUSSION
and Methods) and L-lactate (Biade et al., 1992), are assayed
First, we investigated separately the thermodynamics and enzymatically. Pyruvate is assayed spectrophotometrically
the kinetics of each step of the process. (Biade et al., 1992), and by means of cyclic voltammetry
(Anne et al., 1994). No lactate is produced as long as the
mercury cathode potential does not exceed −1.35 V vs. the
The Keto-Imine Transformation KCl saturated calomel electrode (SCE) on the negative side.
The keto-imine equilibrium (Scheme 1) was characterized At −1.35 V/SCE, both the D and L-alanine concentrations
thermodynamically and kinetically as described previously are 0.16 mM, a result confirming that the reductive amina-
in the case of ␣-keto-glutarate (Anne et al.,1994). tion does take place and produces the racemic D,L-alanine
(concentration: 0.32 mM).
The rate of pyruvate reduction is: dnPRED/dt ⳱ 0.32 ×
10−3 × 20 × 10−3 ⳱ 6.4 × 10−6 mol/h.
It is the imine which is actually reduced. At equilibrium,
in the presence of 1.28M NH3, the imine (CI)e / pyruvate
(CP)e concentration ratio is (CI)e/(CP)e ⳱ K1CNH3 ⳱ 0.256
and (CI)e ⳱ 0.41 mM, (CP)e ⳱ 1.59 mM. The rate of the
imine → keto reaction k−1CIvred is 53 × 10−6 mol/h, i.e.,
much greater than the rate of electrochemical reduction
dnPred/dt. Therefore, the keto/imine transformation is al-
ways at equilibrium in the reduction reactor under such
circumstances and, at constant flow rate and cathode poten-
tial, the rate of electrochemical reduction can be considered
The equilibrium constant K1 ⳱ k1/k−1 can be determined as pseudo-first-order (Bard and Santhanam, 1970; Bard and
by means of cyclic voltammetry, the peak heights being Solon, 1963), i.e., dnPred/dt ⳱ Kred CP with Kred ⳱ 4 × 10−3
proportional to the imine and keto concentrations at equi- L/h. The current i is i ⳱ 2F dnPred/dt and (i/mA) ⳱
librium (Anne et al., 1994). The peak potentials given in 0.216(CP/mM).
Scheme 1 were measured at a potential scan rate of 0.2 V/s As long as the reductive amination is kinetically con-
in the NH3/NH4+ (1.28M / 1.28M) buffer (pH ⳱ 9.2). The trolled by the electrochemical reduction of the imine, the
equilibrium constant and the rate constants k1 and k−1 can be greater the flow rate, the greater Kred. Similarly, the smaller
determined altogether through careful analysis of the cur- the vred/A ratio (A ⳱ area of the mercury pool), the greater
rent vs. time dependence during the electrolysis of the so- Kred. The limiting situation is reached when the rate of
lution at a controlled potential at which only the imine is imine formation becomes the rate-determining step, then
reducible (Anne et al., 1994; Bard and Santhanam, 1970; Kred levels off at Kred,max ⳱ k1vred ⳱ 3.3 × 10−2 L/h. Rais-
Bard and Solon, 1963). We found k1 ⳱ 3.3M−1 h−1, k−1 ⳱ ing vred at constant vred /A ratio would cause an equal in-
16.5 h−1 and K1 ⳱ 0.2. Therefore, in the presence of am- crease in the total volume of the reaction mixture vT which
monia in excess, the ratio of the imine concentration over ought to be kept as small as possible, as discussed later.
the keto concentration is K1CNH3 and the pseudo-first-order
rate constant of imine formation is K1CNH3, CNH3 being the
ammonia concentration in M. The higher CNH3, the more Rate of the Electrochemical Regeneration
of NAD+
favored the imine formation. However, in the inversion pro-
cess, the reaction mixture must be buffered owing to the The flow-through reactor is of the same type as the one
poor stability of NAD+ in exceedingly basic media (Wong already used for the inversion of L-lactate (Biade et al.,

ANNE ET AL.: INVERSION OF CONFIGURATION OF A L-␣-AMINO-ACID 103


1992). The anode is made of carbon felt with an effective 1 to 10 mM range, or at C°A ⳱ 5 mM with C°B in the 5 to
surface area of 5500 cm2, the void volume vox filled with the 50 mM range. Similarly, in the second case, the rate of
solution being 11 mL. To test the regeneration efficiency, a disappearance of Q is measured at C°P ⳱ 10 mM and 0.025
0.5 mM solution of reduced coenzyme is pumped into the ⱕ C°Q ⱕ 0.2 mM or C°Q ⳱ 0.2 mM and 3 ⱕ C°P ⱕ 10
reactor at a flow rate of 100 mL/h. At the output, the en- mM. The respective dependencies of the initial rates Vi2 and
zymatic assay of the solution ascertains that NAD+ is com- Vi−2 of the forward and reverse reactions on C°A and C°B on
pletely regenerated from its reduced forms which can be, in the one hand, and on C°P and C°Q on the other hand, can be
the present case, NADH or the dimers produced by mono- analyzed according to equations (1) and (2).
electronic reduction of NAD+ at a mercury cathode (Biade
et al., 1992). The electrochemical regeneration is as efficient Vi2 Vf
= (1)
in the NH3/NH4+ (1.28M / 1.28M) buffer (pH ⳱ 9.2) as in C°E KMA KMB KiaKMB
1+ + +
other buffers (Biade et al., 1992; Bonnefoy et al., 1988; CA CB CACB
Fassouane et al., 1990; Laval et al., 1987), provided that the
anode potential is positive enough (> 0.4 V/SCE). In the and
following, the rate of electrochemical regeneration of NAD+ Vi–2 Vr
will be always taken as very fast when compared to the = (2)
other key steps of the process. C°E KMP KMQ KiqKMP
1+ + +
CP CQ CPCQ

Enzyme Kinetics Equations (1) and (2) fit well the relevant linear reciprocal
plots of the experimental data. The following quantitative
As previously mentioned, the enzyme catalyzed equilibrium features of the enzyme kinetics can thus be derived: Vf ⳱ 25
(Scheme 2) is thermodynamically unfavorable. and Vr ⳱ 60 ␮mol/min/mg enzyme; they characterize the
enzyme activity in the NH3/NH4+ buffer, and the constants
KMA ⳱ 3.0 mM, KMB ⳱ 12 mM, Kia ⳱ 0.01 mM, KMP ⳱
3.3 mM, KMQ ⳱ 0.12 mM and Kiq < 0.01 mM as well. In the
presence of 10 mM D-alanine, Vi−2 is unaffected and Vi2
undergoes a slight decrease of less than 5%.

In the following, B ⳱ L-alanine, A ⳱ NAD+, P ⳱ py- Simultaneous Occurrence of the Enzymatically


ruvate, Q ⳱ NADH, and Cj ⳱ concentration of species j. Catalyzed and Electrochemically Driven Reactions
The equilibrium constant K2 is defined as: The oxidation and reduction reactors are coupled as previ-
+ ously described for the inversion of L-lactate (Biade et
共CP兲eq共CQ兲eqCNH3CH
K2 = al.,1992). The total volume vT circulating in the compart-
共CB兲eq共CA兲eq ments of the two working electrodes, the reservoir, and the
All the components of the system, including the enzyme, are connecting tubes was minimized at vT ⳱ 38 mL. In a typical
stable in the NH3/NH4+ (1.28M / 1.28M) buffer (pH ⳱ 9.2) experiment, the initial solution contains NAD+ (C°A ⳱ 0.5
over a period exceeding 24 h. mM), L-alanine (C°B ⳱ 10 mM) and L-AlaDH at a con-
At an enzyme concentration of 5 ␮g/mL, the equilibrium centration C°E ⳱ 0.1 mg/mL. A quasi-steady-state is
is reached after 1 h and K2 in the medium is found to be K2 reached after 1 h. Then, the electric currents through both
⳱ 8.0 × 10−14, all concentrations being expressed in M, a reactors have the same absolute value i. Aliquots of the
value of the same order of magnitude as those found in other reaction mixture are taken out at various times t, and D-
buffers (Grimshaw and Cleland, 1981; Yoshida and Freese, alanine is thus enzymatically assayed so as to follow the
1964). The occurrence of the keto/imine equilibrium is progress of its production. The measured D-alanine yield ␳
taken into account for the calculation of the pyruvate con- ⳱ CD−ala/C°B and i vs. t plots are reproduced in Figure 2.
centration at equilibrium in solution. For CH+ ⳱ 10−9.2 M
and CNH3 ⳱ 1.28M, that gives an apparent equilibrium Quantitative Analysis of the Inversion Process
constant K2app ⳱ (CP)eq (CQ)eq /(CB)eq(CA)eq of 1.0 × 10−4.
When the enzyme concentration is lowered, it is possible As the reductive amination is not very efficient the enzy-
to obtain a linear variation of NADH (Q) absorbance at 340 matically catalyzed reaction is essentially driven by the fast
nm over 10 min, starting either with the reactants A and B regeneration of NAD+. Within the oxidation reactor the ki-
(initial concentrations C°A and C°B) to measure the rate of netics is controlled by the enzymatically catalyzed reaction
the forward reaction or with the products P and Q (initial which operates reversibly. Assuming that the enzyme ca-
concentrations C°P and C°Q) to measure the rate of the talysis follows an ordered bibi mechanism, under the pres-
reverse reaction. In the first case, the rate of the appearance ent experimental conditions, its rate VE is given by Segel
of Q is measured at C°B ⳱ 16.7 mM and various C°A in the (1975):

104 BIOTECHNOLOGY AND BIOENGINEERING, VOL. 64, NO. 1, JULY 5, 1999


Taking into account that CQ ⳱ 0, owing to the very fast
electrochemical regeneration of NAD+, and that (Vf /K2app)/
Vr ⳱ (25/60) × 104, equation (3) then reduces to:
VE K2appVrC°ACB
=
C°E KMQCP共1 + C°A Ⲑ Kia兲
Therefore, the rate of production of pyruvate dnPox/dt ⳱
voxVE within the oxidation reactor is:
dnPox K2appVrC°ACB
= v C° × 60 × 10−6
dt KMQCP共1 + C°A Ⲑ Kia兲 ox E
dnPox CB
i.e., = Kox
dt CP
with dnPox/dt in mol/h, va in mL and Vr in ␮mol/min.mg
enzyme, and:

K2appVrC°A voxC°E × 60 × 10−6


Kox = (4)
KMQ 共1 + C°A Ⲑ Kia兲
i.e., Kox ⳱ 3.3 × 10−8 mol/h in the present case.
At quasi-steady-state dnPox/dt ⳱ dnPred/dt, then the pyru-
vate concentration is given by:

Kox
CB
CP
= KredCP and CP = 冑 Kox
Kred
公CB
The rate of electrochemical reduction of pyruvate—dnP /dt
is:
dnP
− = KredCP = 公KredKox 公CB (5)
dt
and

i Ⲑ mA =
2F
3600 冉 冊

dnP
dt
× 1000 = 53.6

× 103 公KredKox 公CB (6)


Figure 2. (a) D-alanine production yield CD-ala/C°L-ala vs.time plot. (b) The rate of production of D-alanine dnD-ala/dt is half that of
Current i vs. time plot (continuous line). C°L-ala ⳱ 10 mM. C°E ⳱ 0.1
mg/mL. C°NAD ⳱ 0.5 mM, flow rate 20 mL/h. Dotted curves ⳱ computed
the electrochemical reduction of pyruvate, thus:
values with Kox ⳱ 3.3 × 10−8 mol/h and Kred ⳱ 4 × 10−3 L/h. (■):
experimental data. The dashed curve in (a) is computed with Kox,max ⳱ 3.3 dnD-ala 1 dnP 公KredKox 公CB
=− =
× 10−7 mol/h and Kred,max = 3.3 × 10−2 L/h (see Text). dt 2 dt 2
or
VE
= dCD-ala 公KredKox 公CB 公KredKox
C°E = −3
= k公CB, with k = ,
dt 2 × 10 vT 2 × 10−3vT

VfVr CACB −
CPCQ
K2app 冊 vT being still expressed in mL. Keeping in mind that C°B ⳱
CB + CD-ala + CP + CI or C°B ⳱ CB + CD-ala + 1.256CP,

Vr KiaKMB + KMBCA + KMACB + CACB +
KMACBCQ
Kiq
because the imine concentration CI is 0.256CP in the pres-
ence of 1.28 M NH3, then:

+
CACBCP
Kip
+冊 Vf

K C + KMPCQ + CPCQ +
K2app MQ P
KMQCACP
Kia or
CB = C°B − CD-ala − 1.256 公Kox Ⲑ Kred 公CB
CB + 2␭公CB − 共C°B − CD-ala兲 = 0

+
CBCPCQ
Kib 冊 (3)
with
and
␭ = 0.628公Kox Ⲑ Kred
公CB = ␭共公1 + 共C°B − CD-ala兲 Ⲑ ␭2 − 1兲 (7)

ANNE ET AL.: INVERSION OF CONFIGURATION OF A L-␣-AMINO-ACID 105


The rate of production of D-alanine becomes: proximations, used for the quantitative analysis of the pro-
cess. Thus, it becomes possible to examine systematically
= k␭共公1 + C°B Ⲑ ␭2 − CD-ala Ⲑ ␭2 − 1 兲
dCD-ala
the effects of the various parameters that can be experimen-
dt tally adjusted so as to improve the productivity, i.e., de-
or crease the time needed to reach a given D-alanine produc-
tion yield ␳ ⳱ CD-ala / C°B. The latter depends on k, ␭, and
d共CD-ala Ⲑ C°B兲 k␭dt C°B, i.e., on Kox, Kred, vT and C°B. Equations (5), (7), and (8)
=
公1 Ⲑ C°B + 1 Ⲑ ␭ 2
− CD-ala Ⲑ C°B␭ − 1 Ⲑ 公C°B
2 公C°B indicate that ␳ would theoretically increase with decreasing
vT and/or increasing Kred and/or Kox. Practically, it is quite
Integration gives: difficult to reduce significantly the total volume vT of cir-
culating solution.
共公1 Ⲑ C°B + 1 Ⲑ ␭2 − 公1 Ⲑ C°B + 1 Ⲑ ␭2 − ␳ Ⲑ ␭2兲 As mentioned earlier, the cathodic efficiency can be en-

冉 冊
hanced by increasing the flow rate and/or by increasing the
1 公1 Ⲑ C°B + 1 Ⲑ ␭2 − 1 Ⲑ 公C°B ratio of the mercury cathode area over the working com-
+ ln
公C°B 公1 Ⲑ C°B + 1 Ⲑ ␭2 − ␳ Ⲑ ␭2 − 1公C°B partment volume inside the reduction reactor. However, as
already pointed out, Kred cannot be raised above the limit
kt Kred,max ⳱ 3.3 × 10−2 L/h which corresponds to the kinetic
= (8)
2␭公C°B control by the rate of imine formation.
According to equation (4), Kox depends on K2app, vox, the
Then equation (8) can be used for the computation of the enzyme kinetic characteristics Vr, KMQ, Kia and the initial
D-alanine production yield ␳ ⳱ CD-ala/C°B vs. t plot. concentrations C°A and C°E. The greater the pH the greater
When CP is negligible compared to C°B −CD-ala, i.e., the apparent equilibrium constant K2app. However, pH can-
practically as long as ␳ ⱕ 0.9, equation (8) reduces to: not be increased because the stability of NAD+ decreases
dramatically above pH 9.2 (Wong and Whitesides, 1981).
公KoxKred KoxKred 2 Once the pH is settled the enzyme kinetic characteristics
␳= t− t (9)
2vT公C°B 16vT2C°B cannot be changed. Any increase in the volume of the work-
ing electrode compartment vox of the oxidation reactor
Computed curves are reproduced in Figure 2a. would cause an equal increase in the total volume of circu-
Equation (6) also becomes: lating solution vT. Due to the low value of Kia, Kox becomes
practically C°A independent once C°A ⱖ 0.5 mM and cannot
i Ⲑ mA = 53.6 × 103␭公KredKoxC°B
be improved by raising C°A above this level. Indeed, the
共公1 Ⲑ C°B + 1 Ⲑ ␭2 − ␳ Ⲑ ␭2 − 1 Ⲑ 公C°B兲 same i is measured at C°A ⳱ 1 mM. Therefore, the only
means of increasing Kox significantly is to raise C°E. We
Equation (8) gives t for each ␳ and the i vs. t plot can also checked out that i undergoes a twofold increase when C°E is
be computed as shown in Figure 2b. increased four times. Practically C°E can be increased 10
times, i.e. up to C° E ⳱ 1 mg/mL, that gives Kox ⳱ Kox,max
⳱ 3.3 × 10−7 mol/h.
Optimization of the Process
The dependence of ␳ ⳱ CD-ala/C°B on C°B is given by
In Figure 2, the dotted curves which fit the measurements equation (8) and more explicitly by equation (9). Clearly,
were computed with the Kox and Kred values corresponding at any time, the greater C°B the smaller ␳. If the times are
to the chosen experimental conditions, namely the initial not too long, ␳ and CD-ala are even proportional to 1/√C°B
concentrations C°E, C°A, C°B, the flow rate, the reactor’s and √C°B respectively. Therefore, if raising CD-ala is the
characteristics, and the electrodes potentials. In the present main purpose, it can be achieved by increasing C°B. The
case, Kox and Kred are 3.3 × 10−8 mol/h and 4 × 10−3 L/h, opposite conclusion holds if raising ␳ is the main purpose.
respectively. Keeping C°B ⳱ 10 mM, the dash-dotted curve in Figure
Figure 2 shows that there is very good agreement be- 2a shows the ␳ growth that could be expected if the experi-
tween the observed and simulated behaviors. Such an agree- mental conditions, allowing for Kox ⳱ Kox,max and Kred ⳱
ment, which holds over a period of 20 d, ascertains the great Kred,max, were fulfilled. Complete inversion of configuration
stability of both the enzyme and the coenzyme in the NH3/ would then be obtained only after 140 h, quite a long time.
NH4+ (1.28 M / 1.28 M) buffer (pH ⳱ 9.2). Taking into Obviously, even under the optimal conditions, the produc-
account that the L-AlaDH molar weight is 228,000 (Grim- tivity remains too low. Therefore, the present process can-
shaw and Cleland, 1981; Yoshida and Freese, 1964), and not be suitable for general use on a preparative scale.
that the enzymatic system undergoes two redox cycles each
time one molecule of D-alanine is produced, the enzyme
CONCLUSION
turnover exceeds 3 × 107.
The remarkable fit between the measured and computed The L → D inversion of an ␣-amino-acid can be carried out
data also justifies the approach, and the accompanying ap- through the combination of the following electrochemically

106 BIOTECHNOLOGY AND BIOENGINEERING, VOL. 64, NO. 1, JULY 5, 1999


driven reactions: selective oxidation of the L-enantiomer by Grimshaw CE , Cleland WW. 1981. Kinetic mechanism of Bacillus subtilis
NAD+ in the presence of the L-␣-amino-acid dehydroge- L-alanine dehydrogenase. Biochem 20:5650–5655.
nase catalyst, electrochemical regeneration of the NAD+ Hilt G, Jarbawi T, Heineman WR, Steckhan E. 1997. An analytical study
of the redox behavior of 1,10-phenanthroline-5,6-dione, its transition-
coenzyme at a carbon felt anode, and reductive amination of metal complexes, and its N-monomethylated derivative with regard to
the ␣-keto-acid at a mercury cathode, the buffer of the re- their as mediators of NAD(P)+ regeneration. Chem Eur J 3:79–88.
action mixture being made of concentrated ammonium/ Hilt G, Lewall B, Montero G, Utley JHP, Steckhan E. 1997. Efficient
ammonia (1.28M / 1.28M). The dehydrogenase exhibits as- in-situ redox catalytic NAD(P)+ regeneration in enzymatic synthesis
tonishing activity and stability under such extreme condi- using transition-metal complexes of 1,10-phenanthroline-5,6-dione
tions of pH and ionic strength; both remain apparently and its N-monomethylated derivative as catalyst. Liebigs Ann /Recueil
2289–2296.
unchanged after 20 d of functioning at room temperature.
Jeffery EA , Meisters A. 1978. Electrochemical synthesis of amino acids by
A careful and detailed quantitative analysis of each of the reductive amination of keto acids. I - Reduction at mercury electrodes.
key steps involved shows that the enzyme catalyzed oxida- Aust J Chem 31:73–78.
tion of the L-enantiomer is so thermodynamically uphill that Jones JB. 1985. Enzymes as chiral catalysts. In: Morrison JD, editor.
it can be driven efficiently to completion only when both the Asymmetric synthesis, chiral catalysts, Vol. 5. New York: Academic
coenzyme regeneration and the ␣-keto-acid reduction are Press. p 309–339.
very effective. The results confirm that the electrochemical Keita B, Essaadi K, Nadjo L, Constant R, Justum Y. 1996. Oxidation
regeneration of NAD+ can be made as fast and efficient as kinetics of NADH by heteropolyanions. J Electroanal Chem 404:
271–279.
required. However, the reduction of the ␣-keto-acid must
Keita B, Essaadi K, Nadjo L, Desmadril M. 1995. Rate-limiting one-
produce only the ␣-amino-acid, i.e., it is the reduction of the electron transfer in the oxidation of NADH by polyoxometalates.
imine that must take place selectively. Then, under the best Chem Phys Lett 237: 411–418.
conditions, the whole process becomes kinetically con- Laval JM , Bourdillon C, Moiroux J. 1987. The electrochemical regenera-
trolled by the rate of imine formation whatever the nature of tion of NAD+ revisited. Biotechnol Bioeng 30:157–159.
the reducing reagent, even if it is a mercury cathode held at Lee LG , Whitesides GM. 1985. Enzyme-catalyzed organic synthesis: A
a finely tuned potential as is the case in the present work. comparison of strategies for in-situ regeneration of NAD from NADH.
Consequently, the process is irremediably slow. At best, the J Am Chem Soc 107:6999–7008.
complete inversion of configuration of a 10 mM solution of Lortie R, Fassouane A, Laval JM , Bourdillon C. 1992. Displacement of
equilibrium in electroenzymatic reactor for acetaldehyde production
L-alanine would require 140 h.
using yeast alcohol dehydrogenase. Biotechnol Bioeng 39:157–163.
Lund H. 1970. Electrochemistry of the carbon-nitrogen double bond. In:
References Patai S, editor. The chemistry of the carbon-nitrogen double bond.
Anne A, Daninos S, Moiroux J, Bourdillon C. 1994. Electrochemical re- New York: Wiley. p 536.
duction of ␣-ketoglutarate in the presence of ammonia as a means of Palmore GTR , Bertschy H, Bergens SH, Whitesides GM. 1998. A metha-
achieving selectively the reductive amination to glutamate. Thermo- nol/dioxygen biofuel cell that uses NAD+-dependent dehydrogenases
dynamic and kinetic characteristics of the keto/imine equilibrium. New as catalysts: Application of an electro-enzymatic method to regenerate
J Chem 18:1169–1174. nicotinamide adenine dinucleotide at low overpotentials. J Electroanal
Bard AJ, Santhanam KSV. 1970. Application of controlled potential cou- Chem 443:155–161.
lometry to the study of electrode reactions. In: Bard AJ, editor. Elec- Persson B, Gorton L, Johansson G, Torstensson A. 1985. Biofuel anode
troanalytical chemistry, Vol. 4. New York: Marcel Dekker. p 215–232. based on D-glucose dehydrogenase, nicotinamide adenine dinucleotide
Bard AJ, Solon E. 1963. Secondary reactions in controlled potential cou- and a modified electrode. Enzyme Microb Technol 7:549–552.
lometry. III - Preceeding and simultaneous chemical reactions. J Phys
Sadakane M, Steckhan E. 1998. Electrochemical properties of polyoxo-
Chem 67:2326–2330.
metalates as electrocatalysts. Chem Rev 98:219–237.
Biade AE, Bourdillon C, Laval JM, Mairesse G, Moiroux J. 1992. Com-
plete conversion of L-lactate into D-lactate. A generic approach in- Segel IH. 1975. Enzyme kinetics. New York: Wiley. p 560–565.
volving enzymatic catalysis, electrochemical oxidation of NADH, and Somers WAC, van Hartingsveldt W, Stigter ECA, van der Lugt JP. 1997.
electrochemical reduction of pyruvate. J Am Chem Soc 114:893–897. Electrochemical regeneration of redox enzymes for continuous use in
Blankespoor RL, Miller LL. 1984. Electrochemical oxidation of NADH, preparative processes. TIBTECH 15:495–500.
kinetic control by inhibition and surface coating. J Electroanal Chem Steckhan E. 1994. Electroenzymatic synthesis. In: Steckhan E, editor. Top-
171:231–241. ics in current chemistry, Electrochemistry V, Vol. 170. Berlin:
Bonnefoy J, Moiroux J, Laval JM, Bourdillon C. 1988. Electrochemical Springer-Verlag. p 84–107.
regeneration of NAD+. A new evaluation of its actual yield. J Chem Tse DCS, Kuwana T. 1978. Electrocatalysis of dihydronicotinamide aden-
Soc Faraday Trans 1 84:941–950. osine diphosphate with quinones and modified quinone electrodes.
Cosnier S, Fontecave M, Innocent C, Nivière V. 1997. An original elec- Anal Chem 50:1315–1318.
troenzymatic system: flavin reductase-riboflavin for the improvement
Williamson DH. 1985. L-alanine. In: Bergmeyer HU, editor. Methods of
of dehydrogenase-based biosensors. Application to the amperometric
enzymatic analysis, Vol. 8, 3rd ed. Verlag Chemie: Weinheim. p
detection of lactate. Electroanalysis 9:685–688.
341–344.
Essaadi K, Keita B, Nadjo L, Constant R. 1994. Oxidation of NADH by
oxometalates. J Electroanal Chem 367:275–278. Wong CH, Whitesides GM. 1981. Enzyme-catalyzed organic synthesis:
Fassouane A, Laval JM , Moiroux J, Bourdillon C. 1990. Electrochemical NAD(P)H cofactor regeneration by using 6-phosphate and the glucose
regeneration of NAD in a plug-flow reactor. Biotechnol Bioeng 35: 6-phosphate dehydrogenase from Leuconostoc mesenteroides. J Am
935–939. Chem Soc 103:4890–4899.
Gra␤l M, Supp M. 1985. D-alanine. In: Bergmeyer HU, editor. Methods of Yoshida A, Freese E. 1964. Purification and chemical characterization of
enzymatic analysis, Vol. 8, 3rd ed. Weinheim: Verlag Chemie. p alanine dehydrogenase of Bacillus subtilis. Biochim Biophys Acta
336–340. 92:33–43.

ANNE ET AL.: INVERSION OF CONFIGURATION OF A L-␣-AMINO-ACID 107

You might also like