You are on page 1of 12

Biotechnology Advances xxx (xxxx) xxx–xxx

Contents lists available at ScienceDirect

Biotechnology Advances
journal homepage: www.elsevier.com/locate/biotechadv

Research review paper

Improved strategies for electrochemical 1,4-NAD(P)H2 regeneration: A new


era of bioreactors for industrial biocatalysis
Clifford S. Morrisona, William B. Armigere, David R. Doddse, Jonathan S. Dordicka,b,c,d,
Mattheos A.G. Koffasa,b,⁎
a
Department of Chemical and Biological Engineering, Rensselaer Polytechnic Institute, 110 8th Street, Troy, NY 12180, United States
b
Department of Biological Sciences, Rensselaer Polytechnic Institute, 110 8th Street, Troy, NY 12180, United States
c
Department of Materials Science and Engineering, Rensselaer Polytechnic Institute, 110 8th Street, Troy, NY 12180, United States
d
Department of Biomedical Engineering, Rensselaer Polytechnic Institute, 110 8th Street, Troy, NY 12180, United States
e
BioChemInsights, Inc., Malvern, PA 19355, United States

A R T I C L E I N F O A B S T R A C T

Keywords: Industrial enzymatic reactions requiring 1,4-NAD(P)H2 to perform redox transformations often require con-
Cofactors voluted coupled enzyme regeneration systems to regenerate 1,4-NAD(P)H2 from NAD(P) and recycle the co-
NADH factor for as many turnovers as possible. Renewed interest in recycling the cofactor via electrochemical means is
NADPH motivated by the low cost of performing electrochemical reactions, easy monitoring of the reaction progress, and
Biocatalysis
straightforward product recovery. However, electrochemical cofactor regeneration methods invariably produce
Electrochemical bioreactors
adventitious reduced cofactor side products which result in unproductive loss of input NAD(P). We review
Cofactor regeneration
Industrial biotechnology various literature strategies for mitigating adventitious product formation by electrochemical cofactor re-
Renalase generation systems, and offer insight as to how a successful electrochemical bioreactor system could be con-
structed to engineer efficient 1,4-NAD(P)H2-dependent enzyme reactions of interest to the industrial biocatalysis
community.

1. Introduction H2. The C-2 and C-6 isomers of the reduced form will be termed 1,2-
NAD(P)H2 and 1,6-NAD(P)H2, respectively, and the more general term
Nicotinamide adenine dinucleotide, commonly written as NAD+/ NAD(P)H2 will be used when specifying a particular isomer of the re-
NADH, and its phosphorylated variants nicotinamide adenine dinu- duced cofactor is not necessary.
cleotide phosphate, NADP+/NADPH, are biochemical redox cofactors Nearly 20% of known oxidoreductases require cofactors to supply
responsible for the transfer of electrons and protons in biological redox stoichiometric quantities of reducing equivalents. For nearly 700
reactions. Formally, this allows the transfer of a molecule of hydrogen known classes of redox enzymes (Ullah et al., 2015), 1,4-NAD(P)H2 is
between metabolic intermediates. The oxidized form of the cofactor the required cofactor. Since the total intracellular pool of both the
accepts electrons, balanced by protons, and in doing so can assume a oxidized and reduced forms of the cofactor is fairly low, the cofactor
number of isomeric structures as well as a dimer. These are shown in undergoes reduction and oxidation turnovers on the order of 103 to 105
Fig. 1. in typical reactions (Angelastro et al., 2017; Ströhle et al., 2016). The
The naming convention for the species in Fig. 1, however, is not current technology for regeneration of 1,4-NAD(P)H2, exemplified in
consistent in the literature, and the common NAD(P)+terminology Fig. 2, includes 1) sacrificial substrates, e.g., a non-valuable alcohol is
causes confusion when considering electrochemical reactions. As ex- present in the reaction mixture and is simultaneously oxidized to an
plained further in Section 2.1, the oxidized form will be termed NAD aldehyde (or acid). This was used by Whitesides and Wong (1982),
(P), and the biologically active reduced form will be termed 1,4-NAD(P) Wong and Whitesides (1983) to utilize all of the reducing power

Abbreviations: NAD, oxidized nicotinamide adenine dinucleotide; NADP, oxidized nicotinamide adenine dinucleotide phosphate; NADH2, reduced nicotinamide adenine dinucleotide
(any isomer); NADPH2, reduced nicotinamide adenine dinucleotide phosphate (any isomer); [NAD]2, nicotinamide adenine dinucleotide dimer; [NADP]2, nicotinamide adenine dinu-
cleotide phosphate dimer; DPN, diphosphopyridine nucleotide; DPNH, reduced diphosphopyridine nucleotide; TPN, triphosphopyridine nucleotide; TPNH, reduced triphosphopyridine
nucleotide; 1,n-NADH2, reduced 1,n-nicotinamide adenine dinucleotide; FAD, flavin adenine dinucleotide; HPLC, high performance liquid chromatography; MSE, saturated mercury
sulfate electrode; mt, metric tonne; SCE, saturated calomel electrode; Poly-His, polyhistidine; HRP, horseradish peroxidase

Corresponding author at: Department of Chemical and Biological Engineering, Rensselaer Polytechnic Institute, 110 8th Street, Troy, NY 12180, United States.
E-mail address: koffam@rpi.edu (M.A.G. Koffas).

http://dx.doi.org/10.1016/j.biotechadv.2017.10.003
Received 23 August 2017; Received in revised form 2 October 2017; Accepted 6 October 2017
0734-9750/ © 2017 Elsevier Inc. All rights reserved.

Please cite this article as: Morrison, C.S., Biotechnology Advances (2017), http://dx.doi.org/10.1016/j.biotechadv.2017.10.003
C.S. Morrison et al. Biotechnology Advances xxx (xxxx) xxx–xxx

Fig. 1. The numbered chemical structures for NAD(P) co-


O NH2
factor material. The isomers of NAD(P)H and a re-
presentative example of an NAD dimer are also shown
H2 N N H
HO O O O- here.
N O P P O N
+ NAD (zwitterion)
N O O O
N

HO OH HO OH

O NH2

H
H2N N -
HO O O O Na+ H
O O
1,4-NADH2
N N P P N
O O O (sodium salt)
N

HO OH HO OH

O NH2

H2 N N H
HO O O O-
O P O + NADP (disodium salt)
N N P N
O O O pKa1 = 3.9; pKa2 = 6.1
N

O OH HO OH

O P O- Na+
O- Na+

O
O NH2 O NH2
NH2
H 3
3 H
H 2
2 4 4 R N 4 4' N R
H
N N 6 5
5
R 1 R 1
6 H2N
H H
O

1,2-NADH2 1,6-NADH2 4,4'-dimer; [NAD]2

H2 N N
HO O O O- Na+
N O P P O
R= N O O O
N
HO OH HO OH

available by employing three enzymes and oxidizing methanol all the Historically, the regeneration of 1,4-NAD(P)H2 by electrochemical
way to CO2; 2) whole cell methods, using resting cells or active fer- means has been well reported, but has never attained wide-spread
mentations, in which a nutrient such as glucose is metabolized to practical application due to the barrier caused by the formation of
provide reducing equivalents (Zaks and Dodds, 1997); and 3) combined biologically inactive isomers of NAD(P)H2 and dimers of NAD(P). The
chemoenzymatic methods (Eikeren, 1989). While all of these methods recognition of these well-known chemistries has held back the appli-
can be used, they require handling whole cells, multiple enzymes, and/ cation of electrochemical 1,4-NAD(P)H2 regeneration to industrial
or other molecules, such as the sacrificial substrate. Direct regeneration processes. As discussed in Section 4.2, strategies for dimer mitigation
of the 1,4-NAD(P)H2 would be simpler and less expensive if electrons have been recognized for decades, but no similar approaches for isomer
and protons could be provided directly to the oxidized form of NAD(P). mitigation were known. However, naturally occurring enzymes (re-
Electrochemical methods, under specific reaction conditions, are able to nalases) were recently discovered, as discussed in Section 4.2.1, in both
achieve exactly that. eukaryotic and prokaryotic cells whose primary function is the recovery
Formally, the reduced form of the cofactor, NAD(P)H2, carries one of the biologically inactive forms of 1,2- and 1,6-NAD(P)H2 by re-oxi-
molecule of H2, which is provided to the reaction being catalyzed by the dizing them back to NAD(P). This recent discovery of nature's system
enzyme requiring reducing power. Reduced cofactor is thus the che- for conserving the biologically active NAD(P) molecule within the cell
mical equivalent of hydrogen, and competition with other sources of can be incorporated into electrochemical systems for removing this
hydrogen must be considered when planning a chemical process that barrier.
could use NAD(P)H2. An electrochemical bioreactor must achieve a number of cofactor

2
C.S. Morrison et al. Biotechnology Advances xxx (xxxx) xxx–xxx

Fig. 2. Examples of current industrial strategies used for


SACRIFICIAL
SUBSTRATE the regeneration of 1,4-NAD(P)H2. 1,4-NAD(P)H2 provides
reducing power to a variety of biocatalytic processes.
reduced cofactor
oxidized 1,4-NAD(P)H2 substrate
by-product
1st 2nd
redox redox
enzyme enzyme

sacrificial
NAD(P) product
substrate
oxidized cofactor

WHOLE CELL
nutrient substrate
e.g. glucose

cell

oxidized metabolite(s) product


e.g. gluconolactone, CO2

CHEMO-ENZYMATIC

chemical reagent reduced cofactor


oxidized by-
1,4-NAD(P)H2 substrate
e.g. mBH4 product
e.g. CH3CHO 1st 2nd
redox redox
enzyme enzyme
spent reagent sacrificial (reduced)
e.g. substrate NAD(P) product
mB(OH)4 e.g. CH3CH2OH oxidized cofactor

turnovers to maximize process efficiency and economic value. To when writing balanced electrochemical reactions showing the formal
maximize the number of cofactor turnovers, strategies for mitigating transfer of a hydrogen molecule, as two electrons and two protons, to
the formation of the adventitious but undesired reduced cofactor side the oxidized species NAD(P). These molecules are best regarded as a
products, 1,2- and 1,6-NAD(P)H2 as well as the dimer [NAD(P)]2, must neutral zwitterion, in the same way that amino acids are regarded, in
be considered. These mitigation strategies can take the form of re- order to account for a holistic perspective on the overall charge balance
covering NAD(P) or 1,4-NAD(P)H2 from the undesired reduction pro- within their native biological environments. To avoid further confusion,
ducts or preventing their formation in the first place. this review suggests to the chemical community the nomenclature of
This review surveys the literature of NAD(P)H2 reduction and oxi- Reaction 2 below, despite the lengthy historical use of the nomen-
dation over the last 70 years and documents the important electro- clature of Reactions 1a and b.
chemical products resulting from NAD(P) reduction along with the In its oxidized state, non-phosphorylated NAD is an electron ac-
observations that have confounded and confused the field for many ceptor with a formal positive charge at the quaternary nitrogen atom of
decades. This literature review, along with recent developments in the the nicotinamide ring and two ionizable hydroxyl groups on the di-
last 18 months, provides insights into the issues that have held back the phosphate linkage. The pKa of these hydroxyl groups are approximately
use of electrochemical regeneration of 1,4-NAD(P)H2. 2.4 and 6.6, respectively. The more acidic hydroxyl provides a formal
negative charge on the molecule to satisfy the positive charge on the
nitrogen atom in the nicotinamide ring as a zwitterion. Depending on
2. Redox enzymes and pyridine nucleotide cofactors
the exact pH, the second hydroxyl group will be deprotonated to
varying degrees giving NAD a net negative charge, contrary to the
2.1. Properties of NAD(P)H2
commonly used NAD + notation. This net negative charge is balanced
by a suitable cation in aqueous solution.
As of this review, thousands of articles in the scientific literature
A similar situation exists with NADP, albeit with two additional
refer to a charged species NAD(P)+ as part of a redox pair with neutral
ionizable hydroxyl groups on the extra phosphate moiety. Thus NADP
NAD(P)H2. Indicating a positive charge on the nicotinamide moiety of
should be properly represented as a di- or even tri-anionic species, not
NAD(P) without the context of the rest of the molecule leads one to
as NADP +, which suggests it is a mono-cation.
conclude that the entire molecule has a formal charge of + 1 due to the
The aforementioned discussion on net charge is important when
nitrogen at the 1-position of the nicotinamide ring. This is commonly
considering electrochemical redox reactions in solution. Critically, two
represented in the literature by either of the following representations:
electrons and two protons are required for the reduction of NAD(P) to
NAD(P)+ + H+ + 2e− → NAD(P)H (a) NAD(P)H2, and this is clear experimentally using mass spectrometry
NAD(P)+ + 2H+ + 2e− → NAD(P)H + H+ (b) Reaction (1) with NAD and NADH2 having a molecular weight difference of 2 amu.
Therefore, it is more accurate to report the overall redox reaction with
While commonly encountered, these descriptions lead to confusion

3
C.S. Morrison et al. Biotechnology Advances xxx (xxxx) xxx–xxx

Fig. 3. The mechanism for the electrochemical production


O NH2 O NH2
of 1,4-NADH2 and [NAD]2 involving one-electron transfer
H reactions with a radical intermediate. The radical species is
- H H+
H the branching point of the reaction.

N Step 2a N
R R
e- 1,4-NADH2
O NH2 O NH2 Step 2

H H
e-
N+ Step 1 N Step 3
O
R R NH2
NAD NAD NAD
R N N R 4,4'-dimer; [NAD]2

H2N
H2N N
HO O O O- Na+ O
N O P P O
R= N O O O
N

HO OH HO OH

the following representation: paper as [NAD(P)]2, is also reported to form during the electrochemical
reduction of NAD(P). This reaction also locks some of the cofactor
NAD(P) + 2H+ + 2e− → NAD(P)H2 Reaction (2)
material into a biologically inactive form with each electrochemically-
In some reports, NAD(P)H2 is referred to in its enzymatically active driven reduction cycle. This problem, referred to as “the nonspecific
form as 1,4-NAD(P)H2 in order to specifically reflect that the hydrogen reduction problem,” was pointed out in the 1980s by Whitesides et al.
on the dihydropyridyl moiety is at the C-4 position. Other reported as a major reason for the inability of electrochemical 1,4-NAD(P)H2
names for the non-phosphorylated 1,4-NADH2 include β-NADH, DPNH, regeneration methods to be competitive with biological methods
and reduced coenzyme I. Other reported names for the phosphorylated (Chenault and Whitesides, 1987).
form, NADPH2, include β-NADPH, TPNH, and reduced coenzyme II. 1,2-NAD(P)H2 has a half-life of ~30 min (Beaupre et al., 2015).
1,4-NAD(P)H2 is significant in a variety of industrial processes. For However, this was probably not known to earlier researchers in the
example, the enzymatic production of butanol (Kang et al., 2007) and field who reported unusual situations where their experiments did not
ethanol (Shin et al., 2004); the hydroxylation of alkanes and aromatic give expected results. For example, Jaegfeldt (1981) observed dis-
compounds (Urlacher and Eiben, 2006); the epoxidation of alkenes crepancies in HPLC analyses of NADH2 isomer preparations. In another
(McClay et al., 2000); heteroatom oxygenations and Baeyer–Villiger study, 1,6-NAD(P)H2 was positively identified by Godtfredsen and Ot-
reactions (Zambianchi et al., 2002); as well as the hydrolysis of tri- tesen as an inhibitor of lactate dehydrogenase (Godtfredsen and
glycerides (Hollmann et al., 2006) are a few of the relevant industrial Ottesen, 1978), but the authors were probably unaware that 1,2-NAD
biocatalytic processes that utilize 1,4-NAD(P)H2 as a redox cofactor. (P)H2 also behaves as an inhibitor of 1,4-NAD(P)H2-dependent enzymes
1,4-NAD(P)H2 must be used in stoichiometric quantities for these as the isomer's instability likely prevented its detection or isolation
biocatalytic reactions, but the high cost and lability (Mädje et al., 2012) (Beaupre et al., 2015). Similarly, Biellman et al. reported difficulties in
of these cofactors presents an imposing economic obstacle for industrial isolating and studying certain inhibitors of lactate dehydrogenase
processes relying on 1,4-NAD(P)H2 without the ability to reduce NAD (Biellmann et al., 1979).
back to 1,4-NAD(P)H2. Furthermore, it is disadvantageous to add high The mechanism for the electrochemical formation of the 4,4′-dimer
concentrations of cofactor to an in vitro enzyme reaction since excess [NAD]2 is shown in Fig. 3. A one-electron reduction occurs at a re-
amounts can lead to enzyme inhibition (Williams, 1952). Therefore, ducible position of the pyridine ring to form an NAD% radical, which
there has been motivation to achieve continuous in situ regeneration of then undergoes either a second electron transfer to form an anion
1,4-NAD(P)H2 from a limited pool of NAD(P). (which is then quenched by a proton) or a radical dimerization to form
Several approaches to 1,4-NAD(P)H2 regeneration have been in- an [NAD]2 dimer. If the second electron transfer is slower than di-
vestigated, including electrochemical, photochemical, and chemical merization, then dimer will be the predominant product (Ali et al.,
reduction methods (Lee and Whitesides, 1985) in addition to whole cell 2012b). Migration of the radical leads to the 1,2-NADH2 or the 1,6-
and coupled enzyme/substrate methods (Zhao and van der Donk, NADH2 isomers.
2003). Due to their low cost and simplicity, electrochemical techniques Given the three reducible positions on the pyridine ring, as well as
have received increased attention, and therefore this review focuses on additional stereoisomers of the dimer due to the asymmetric carbons at
1,4-NAD(P)H2 regeneration in electrochemical systems. the point of dimerization, one might expect a combinatorial number of
possible dimers, e.g. 2,2′-, 2,4′-, 2,6′-, 4,4′-, 4,6′-, and 6,6′-dimers, plus
2.2. The issue of nonspecific NAD(P) reduction the stereoisomers for each dimer, thus totaling 21 possible dimers. The
dimers are typically named by the combination of monomeric species
As opposed to enzymatic regeneration methods, which are highly that constitute the dimer, e.g. n,m′-[NAD(P)]2.
selective for the catalytically active form 1,4-NAD(P)H2, non-biological Experimental evidence suggests that only seven of the 21 possible
regeneration strategies are driven by nonspecific electrochemical, dimers are sufficiently stable to persist in solution. These are the three
photochemical or chemical reduction (Lee and Whitesides, 1985). Such possible 4,4′-dimers and the four possible 4,6′-dimers. Indeed, three
nonspecific approaches result in a mixture of reduced nicotinamide ring stereoisomers of 4,4′-dimers comprise 90% of electrochemically pre-
isomers, including the desired 1,4-species and the non-biologically ac- pared dimer solutions, while the other 10% consists of three stereo-
tive 1,2- and 1,6- reduction isomers (Chenault and Whitesides, 1987). isomers of 4,6′-dimers. No 6,6′-dimers nor any dimers involving the C-2
The formation of the inactive isomers leads to inefficiency in co- position were found (Jaegfeldt, 1981). Generally similar results were
factor regeneration systems and ultimately limits the recycling turnover obtained by Carelli et al., who observed only 4,4′-dimers after
due to dead-end products. In addition, a dimer of NAD(P), noted in this

4
C.S. Morrison et al. Biotechnology Advances xxx (xxxx) xxx–xxx

externally driven current flow commercially feasible. Moreover, the low surface area of these elec-
trodes results in a slow NAD(P) conversion rate, and the poorly-defined
fluid flow and mass transport characteristics of laboratory-scale elec-
oxidized trochemical systems present scale-up challenges (Walsh and Reade,
cofactor
1994). In addition, high overpotentials or use of electron mediators are
Anode Cathode often required for the direct reduction of NAD(P) to 1,4-NAD(P)H2 (Lau
6 H2O reduced et al., 2005). Electrochemical bioreactor technology development for
4 e- 4 e- cofactor
research use or for industrial production should, therefore, focus on
systems capable of achieving a practical number of cofactor turnovers
O2
in a reasonable amount of time.
+
The turnover efficiency for cofactors regenerated by any strategy
4 H3O+ 4 H3O+ 4 H2O
must be high enough to render the adventitious formation of useless
reduced cofactor products, i.e. the loss of usable cofactor per each
Anode Chamber Cathode Chamber turnover, negligible. Regenerating the reduced form of cofactor in situ
allows for large-scale reactions to proceed economically. Though to be
proton-
permeable
truly economical, the regeneration method should be capable of re-
membrane cycling the cofactor 102–106 times, depending on the cost of the co-
factor and the value of the products made by the process. The turnover
Fig. 4. The overall electrochemistry and general arrangement of the electrochemical cell.
Shown here is the direct electrochemical reduction of the oxidized form of the cofactor to efficiency required to have at least 50% remaining cofactor activity
the reduced form of the cofactor. after 102 turnovers is 99.3% and for 106 turnovers the efficiency needed
is 99.99993%. It then follows that of the various regeneration strate-
gies, the only strategy historically capable of delivering such turnover
electrochemical reduction (Carelli et al., 1981), and Kovář et al. who
efficiencies are those mediated by coupled redox enzyme reactions
found that 4,4′-dimers constituted the majority of electrochemical
using sacrificial substrates (Chenault et al., 1988).
dimer preparations, with only small amounts of 4,6′-dimers present in
Coupled redox enzyme reactions, whether through the use of sa-
the mixtures (Kovář et al., 1985). Thus, dimers of the 4,4′ type appear
crificial substrates or resting whole cell methods, utilize the simulta-
to form the most stable structures out of the 21 possible dimers, as
neous reduction of NAD(P) in one direction and the oxidation of 1,4-
predicted by Burnett and Underwood (Burnett and Underwood, 1968).
NAD(P)H2 in the other direction. This can occur by using an enzyme
with a reversible reaction in both reaction directions but acting on
3. Electrochemical bioreactors different substrates, or by using two enzymes with different activities
for different substrates that, when combined, result in the cycling of
3.1. Function of electrochemical bioreactors NAD(P)/1,4-NAD(P)H2 (see Fig. 2). Indeed, the production of many
modern industrially-relevant biocatalysis products requiring the re-
Electrochemical bioreactors operate on the principle of either pro- generation of 1,4-NAD(P)H2 are only made possible by the use of
ducing or using electrical current. Examples of electrochemical bior- coupled enzyme systems. Examples of such products include the statin
eactors include bio-batteries, microbial fuel cells, microbial electro- drug Lipitor® (atorvastatin calcium) (see Fig. 5) (Schroer and Lütz,
synthesis cells, enzymatic electrolysis cells, and electrolyzers (Rabaey 2009; Schroer et al., 2007; Ma et al., 2010), the antilipemic agent
et al., 2010). Electrolyzers are used most frequently in cofactor re- ZETIA® (ezetimibe) (Homann and Previte, 1996), the HIV protease in-
generation and are thus the focus of electrochemical bioreactors de- hibitor REYATAZ® (atazanavir) (Patel, 2006), the anti-malarial drug
scribed in this review. precursor amorphadiene (Ma et al., 2011), and the pharmaceutical
As shown in Fig. 4, the electrolyzer consists of an anode chamber, building block L-tert-leucine (Leuchtenberger et al., 2005) among other
where four electrons are obtained from O2 reduction to water, and a drugs (Zaks and Dodds, 1997).
cathode chamber, where these electrons are donated to the oxidized Electrochemical cofactor regeneration falls broadly into two stra-
cofactor to generate the reduced cofactor. The electrode chambers are tegies: 1) direct reduction of NAD(P) in the cathode chamber, and 2)
typically separated by an ion-selective membrane to allow for the indirect reduction of NAD(P) using a mobile electron transport med-
transfer of protons in the form of hydronium ions to cross from the iator in the cathode chamber that is recycled between an enzyme re-
anode chamber to the cathode chamber while the electrons are trans- action chamber where NAD(P) reduction takes place with the oxidized
ferred from the anode to the cathode chamber over the wire connecting electron mediator returning to the cathode chamber. In general, elec-
the chambers. tron transport mediators can be scaled to maximize the availability of
Reduced cofactor can be regenerated electrochemically with as electron-donating species throughout the system. This could help to
much as 98% of 1,4-NAD(P)H2 regenerated from NAD(P) in solution. minimize electrolyzer size relative to reactor volume.
This highly efficient regeneration of 1,4-NAD(P)H2 was reportedly ob- An electrochemical bioreactor system capable of simultaneously
tained with an electrochemical bioreactor utilizing a glassy carbon performing an 1,4-NAD(P)H2-dependent reaction of interest while ef-
electrode as the cathode (Ali et al., 2012b; Ali et al., 2011). While this is ficiently regenerating 1,4-NAD(P)H2 in situ offers an advantage over
promising, most systems reported to date are too small to be currently-practiced technology by eliminating the requirements of

substrate Fig. 5. An example of a coupled enzyme system for the


regeneration of 1,4-NADPH2. 1,4-NAD(P)H2 provides re-
reduced cofactor O O
ducing power to a variety of industrially-relevant products
oxidized by-product 1,4-NADPH2 Cl
O such as pharmaceutical intermediates, adapted from (Ma
gluconate, C6H10O6
et al., 2010).
GDH KRED

OH O
sacrificial (reduced)
NADP Atorvastatin
substrate Cl
glucose, C6H12O6 oxidized cofactor O
reduced product

5
C.S. Morrison et al. Biotechnology Advances xxx (xxxx) xxx–xxx

Table 1 number of NAD turnovers achieved (Hollmann et al., 2001).


Five guiding principles for process development for a successful electrochemical bior- Some of the more successful embodiments of 1,4-NAD(P)H2 re-
eactor.
generating systems utilize glassy carbon electrodes and their modified
Principle Function counterparts to achieve efficient regeneration. Azem et al. reported in
2004 that a glassy carbon electrode modified with a submonolayer of
Reduction Reduction of oxidized form of the cofactor ruthenium is capable of the direct reduction of NAD to NADH2 with
Reaction Useful reduction of substrate to desired product with concomitant
little to no concomitant formation of [NAD]2 or 1,6-NADH2. The au-
oxidation of the cofactor
Recovery Recovery of unused forms of reduced cofactor (i.e. non- thors made no mention of 1,2-NADH2 in their system. Overall, the ru-
biologically useful isomers, dimers, etc.) thenium modified glassy carbon electrode allowed a 96% yield of en-
Regeneration Reduction of recovered oxidized cofactor zymatically-active 1,4-NADH2. They reported that the reaction was
Recycle Practical engineering to allow continuous closed system cofactor highly irreversible, independent of pH, and occurs at high cathodic
recycle
overpotentials where the reaction is diffusion controlled and is first
order with respect to NAD. The authors postulated that ruthenium
using a sacrificial substrate, which requires separation of its associated serves simultaneously to provide sites for adsorbed hydrogen on the
products from the desired product. electrode, thereby selectively protonating NAD(P)% radicals as they
form, as well as acting as a physical barrier against the formation of
dimers at the electrode surface (Man and Omanovic, 2004; Azem et al.,
3.2. Criteria for successful electrochemical bioreactors 2004).
Ali et al. observed that even unmodified glassy carbon electrodes
The co-production of 1,2- and 1,6-NAD(P)H2 and [NAD(P)]2 re- are capable of selective 1,4-NADH2 regeneration. At −1.40 V vs. MSE
mains a challenge for electrochemical bioreactor systems. Therefore, on an unmodified glassy carbon electrode, the yield of 1,4-NADH2 was
successful electrochemical bioreactor systems need to follow five only ~32%. However, at a high overpotential of − 2.30 V vs. MSE on
guiding principles for process development, as shown in Table 1. the same electrode, the yield of 1,4-NADH2 was ~98% (Ali et al.,
Recovery processes, which have yet to be given proper treatment in 2011). Various modifications to glassy carbon electrodes have been
the literature in the context of electrochemical bioreactor systems, are explored to lower high overpotentials. For example, Ali et al. reported
vital to the success of a process by which 1,4-NAD(P)H2 is regenerated that their glassy carbon electrode nano-patterned with platinum and
electrochemically. Therefore, it is imperative for the success of an nickel gave 100% yield of 1,4-NADH2 from NAD. They posited that
electrochemical bioreactor process to follow the above-mentioned platinum and nickel on the electrode surface provided sites for ad-
guidelines and to include “recovery” as a fifth necessary principle. It is sorbed hydrogen close to the sites of dimer formation in order to in-
essential that NAD(P) is recovered from the adventitious reduced co- crease radical protonation kinetics. The authors state that this electrode
factor products, namely 1,2- and 1,6-NAD(P)H2 and [NAD(P)]2, so that has good potential for industrial applications since they were able to
recycling the cofactor results in a minimal loss of cofactor per turnover operate at relatively lower overpotentials than on an unmodified glassy
event, and the overall process can be performed as economically as carbon electrode (Ali et al., 2012a).
possible. These authors also indicated other electrode configurations could be
used. Specifically, a carbon nanofiber cathode grown on a stainless steel
3.3. Electrochemical 1,4-NAD(P)H2 regeneration mesh was able to yield ~99.3% 1,4-NADH2 at −2.3 V vs MSE. The
authors posited that this electrode could be more commercially viable if
Various embodiments of electrochemical 1,4-NAD(P)H2-re- the carbon nanofiber cathode is nano-patterned with platinum or nickel
generating systems have been reported in the literature to varying de- and further suggested that the high overpotentials could be reduced to
grees of success. Elving et al. reported in 1982 that the direct reduction − 1.5 V vs. MSE with the nano-patterned platinum or nickel (Ali et al.,
of NAD(P) on an unmodified metallic electrode generally results in the 2012b). Ullah et al. reported that a titanium substrate coated with an
formation of [NAD(P)]2 via NAD(P)% radicals which can be reduced to iridium/ruthenium-oxide coating gave a yield of ~88% 1,4-NADH2 at
1,4-NAD(P)H2 and its isomers at high overpotentials (Elving et al., − 1.70 V vs. MSE. The authors attribute this result to the relatively
1982). Since operating at higher overpotentials is not economically weak hydrogen-metal bonds formed on the electrode surface which
efficient, much work in the literature has focused since then on creating provides surface-adsorbed hydrogen to NAD(P)% radicals to favor 1,4-
modified electrodes to promote 1,4-NAD(P)H2 formation at lower NAD(P)H2 formation rather than dimer formation while requiring less
overpotentials. energy to break the bond from the electrode surface, resulting in a
Siu et al. reported the conversion of α-ketoglutarate to glutamate by lower overpotential (Ullah et al., 2015). This result is particularly in-
glutamate dehydrogenases following the direct electrochemical reduc- teresting in that the authors achieved a relatively high yield of 1,4-
tion of NAD in the presence of vanadia-silica xerogels using unmodified NADH2 at a relatively low overpotential without using a glassy carbon
platinum electrodes. The authors reported that 3300 turnovers of NAD electrode.
were achieved using their system, but they do not discuss formation of Although many of these reports state there is little to no formation
adventitious side products that may have limited the total number of of 1,6-NAD(P)H2, 1,2-NAD(P)H2, or [NAD(P)]2, these experiments were
turnovers (Siu et al., 2007). done at a relatively small scale compared to what would be expected for
Hildebrand et al. reported the conversion of acetophenone to (R)- industrial reactors. It may be reasonable, therefore, to expect greater
phenylethanol by alcohol dehydrogenase from Lactobacillus brevis fol- variability in 1,4-NAD(P)H2 regeneration efficiencies during scale-up.
lowing the indirect reduction of NAD on a carbon felt electrode using To that end, Wang and Yiu demonstrated for the first time the use of
Cp*Rh(bpy) as an electron mediator. The authors focused on enzyme a heterogeneous catalyst with H2 as a reductant to regenerate 1,4-
productivity rather than cofactor turnover numbers, so their system NADH2 in situ. The yield of 1,4-NADH2 was found to increase with
with 35 turnovers of NAD does not immediately lend itself to industrial temperature and pH. The 1,4-NADH2 regeneration was coupled with
application (Hildebrand and Lütz, 2007). the conversion of propanal to propanol by alcohol dehydrogenase. The
Hollman et al. reported the conversion of several α-substituted turnover number for the cofactor was found to be 7. Interestingly, the
phenol derivatives to their ortho-hydroxylated counterparts by 2-hy- catalyst appears to provide surface reactive hydrogen to reduce NAD to
droxybiphenyl-3-monooxygenase following the indirect electro- 1,4-NADH2 (Wang and Yiu, 2016). Although the mechanistic details are
chemical reduction of NAD on a glassy carbon electrode using [Cp*Rh unclear, such surface reactive hydrogen is analogous to the findings of
(bpy)Cl]Cl as an electron mediator. No information was provided on the Azem, Man and Omanovic, Ullah, and Ali. This is indicative of a more

6
C.S. Morrison et al. Biotechnology Advances xxx (xxxx) xxx–xxx

fundamental, underlying mechanism involving chemisorbed hydrogen 1,4-NAD(P)H2.


for selective 1,4-NAD(P)H2 regeneration irrespective of the regenera- Clearly, operational controls and conditions can be implemented
tion driving force. into engineered electrochemical bioreactor systems to favor the pro-
duction of the catalytically active 1,4-NAD(P)H2 isomer while si-
4. Mitigation strategies multaneously discouraging the formation or persistence of unfavorable
isomers and dimers. Optimization of any particular electrochemical
4.1. Prevention bioreactor process should begin by investigating the optimum buffer
pH, buffer composition, electrolysis potential, and various other op-
One of the most significant factors in achieving efficient electro- erational controls for the process of interest. In the likely case that there
chemical regeneration of 1,4-NAD(P)H2 is the choice of electrode ma- is no optimal set of conditions under which to selectively form 1,4-NAD
terials. In general, the electrode should be engineered to improve the (P)H2, recovery processes should be implemented in tandem with pre-
kinetics of the second electron transfer (Step 2) versus the dimerization vention strategies in order to avoid wastage of NAD(P).
of two NAD(P)% radicals (Step 3) in Fig. 3 or to prevent dimerization
altogether. This can be achieved by physically preventing the coupling 4.2. Recovery
of NAD(P)% radicals on the electrode surface, or by enhancing hydrogen
and electron transfer kinetics so that the rate of formation of all isomers 4.2.1. Enzymatic recovery of NAD(P) cofactor material by renalase
of NADH2 outcompetes the rate of dimer formation. In 2005, Xu et al. reported the discovery of a novel enzyme, called
Long and Chen reported that a silver electrode modified with L- renalase, secreted into the blood by the human kidney that modulates
histidine resulted in the selective formation of 1,4-NADH2 without systemic blood pressure by metabolizing circulating catecholamines.
concomitant formation of [NAD]2. They used an alcohol dehydrogenase While the highest renalase gene expression was found in the kidney, the
to catalyze the conversion of acetaldehyde to ethanol with an 82% yield enzyme was also expressed in cardiac tissue, skeletal muscle tissue, and
of active 1,4-NADH2 and a cofactor turnover number of 9 (Long and the small intestine. The enzyme was characterized as an FAD-dependent
Chen, 1997). Although this turnover number is commercially im- amine oxidase important to human health (Xu et al., 2005). In 2010,
practical, it is nonetheless interesting that relatively high selectivity Pandini et al. reported that by expressing recombinant human renalase
was achieved in the formation of 1,4-NADH2. in E. coli with an N-terminal poly-His tag, enough enzyme could be
Along similar lines, Baik et al. reported that a yield of ~10% 1,4- obtained to perform biochemical studies (Pandini et al., 2010).
NADH2 was achieved on a bare gold amalgam electrode, but a yield of Beaupre et al. reported that there had been no convincing demon-
~ 75% 1,4-NADH2 was achieved on a gold amalgam electrode modified stration of the claim that renalase does in fact catalyze the oxidation of
with cholesterol. The authors speculated that cholesterol likely served circulating catecholamines since earlier work did not sufficiently con-
as a physical barrier to the formation of [NAD]2 on the electrode sur- trol for spontaneous oxidation of catecholamines in buffers containing
face. Using this system, they were able to catalyze the conversion of dissolved oxygen. Hence, the physiologically-relevant metabolic func-
pyruvate to D-lactate by D-lactate dehydrogenase. The NAD turnover tion of the enzyme remained unclear (Beaupre et al., 2013). In later
number was reported to be 1400. The authors also suggested that work, this group reported that renalase catalyzes the oxidation of the
keeping the concentration of NAD low discourages [NAD]2 formation undesired isomeric forms of NAD(P)H2 that may arise in vivo via events
and increases total NAD turnovers since there would be a smaller such as tautomerization and nonspecific redox events. They postulated
probability of NAD% radicals undergoing dimerization if there are less that the biological significance of such an enzyme could be to act as a
radicals available in the first place (Baik et al., 1999), and this will be housekeeping enzyme that prevents 1,2- or 1,6-NAD(P)H2 from in-
clear from inspection of Fig. 3 above. hibiting 1,4-NAD(P)H2-dependent reactions, as they had found that 1,2-
As mentioned in Section 3.3, it has been demonstrated that bare and and 1,6-NAD(P)H2 is highly inhibitory to dehydrogenase enzymes. The
modified glassy carbon electrodes are also capable of the selective biochemical reaction catalyzed by renalase is given in Fig. 6. Renalase
formation of enzymatically active 1,4-NAD(P)H2 via electrolysis. This requires an oxidizing environment to catalyze the reaction in which
selectivity is likely due to surface adsorbed hydrogen, which is avail- 1,2- and 1,6-NAD(P)H2 are oxidized and the two electrons and two
able to hydrogenate NAD% radicals in apparently the correct orientation protons are eventually passed on to O2 as an electron acceptor to form
to minimize the formation of isomers and dimers. Various catalysts can H2O2 (Beaupre et al., 2015).
also be employed to encourage the favored reactions. Interestingly, a bacterial analog of renalase has been identified in
Established electrochemical preparations of [NAD(P)]2 use alkaline Pseudomonas phaseolicola which features the same substrate specificity
buffers to stabilize the dimer for further use or analysis (Jaegfeldt, and catalytic function as human renalase. Hoag et al. reported that this
1981; Schmakel et al., 1975). The dimer is known to be acid-labile bacterial renalase is characterized by a marked substrate preference for
(Schmakel et al., 1975; Bresnahan and Elving, 1981), so 1,4-NAD(P)H2 NAD isomers over of NADP isomers. The existence of a renalase analog
regeneration systems seeking to avoid the formation of dimer may in an organism without a circulatory system lends further credibility to
benefit from the use of a slightly acidic buffer if the prepared NAD(P)H2 the hypothesis that renalase is a cellular housekeeping enzyme (Hoag
is to be used quickly after preparation. et al., 2015) and not a modulator of blood pressure as was originally
Indeed, it has been reported that NAD(P)H2 can be electro- proposed.
chemically prepared in higher concentrations relative to other mixture Recent fundamental work by Moran et al. has been convincing in
components by increasing the concentration of protons in the buffer elucidating the catalytic activity of renalase, and by extension, the
(Jaegfeldt, 1981). potential biological implications of the various NAD(P)H2 isomers with
Azem et al. reported an apparent voltage selectivity for the amount respect to primary metabolism (Moran, 2016). The potential implica-
of enzymatically active NADH2 prepared electrochemically at various tions renalase has for applied biochemistry and industrial applications
electrolysis potentials (Azem et al., 2004). This was corroborated in the are the subject of patent applications (Morrison et al., 2017). Since the
reports by Ullah et al. and Ali et al. where the amount of 1,4-NADH2 catalytic activity of renalase is fundamentally a recovery reaction, an
produced via electrolysis was maximized at a specific electrolysis po- engineered electrochemical bioreactor system taking advantage of im-
tential relative to the choice of electrode materials and running buffers mobilized renalase could theoretically ameliorate the adventitious
(Ullah et al., 2015; Ali et al., 2011; Ali et al., 2012a). This clearly in- electrochemical production of the undesired isomers of NAD(P)H2 that
dicates that the electrochemical properties of 1,4-NAD(P)H2 allow for a would otherwise inhibit 1,4-NAD(P)H2-dependent enzymes that gen-
voltage selectivity effect that an engineered electrochemical bioreactor erate the product of interest, and potentially lock all NAD(P) cofactor
system could take advantage of in order to maximize the production of material into an unusable form. Either of these events severely limit the

7
C.S. Morrison et al. Biotechnology Advances xxx (xxxx) xxx–xxx

H2O2 H2O2 Fig. 6. Renalase catalyzes an NAD(P) recovery reaction.


O NH2 O2 O NH2 O2 O NH2
The biochemical reaction catalyzed by renalase oxidizes
H
H 1,2- and 1,6-NAD(P)H2 to NAD(P) with concomitant pro-
H H
H
duction of H2O2.
N+
renalase renalase
N N
R R R
H H

1,2-NAD(P)H2 NAD(P) 1,6-NAD(P)H2

H2N N -
HO O O O
N O P P O
N O O O
R= N
HO OH HO OH

economic potential of any industrial process relying on the electro- acceptor in this reaction (Avigliano et al., 1983). 1,4-NADH2 is also
chemical regeneration of 1,4-NAD(P)H2. known to participate in photooxidation reactions, presumably through
similar mechanisms as those proposed by Noguchi et al. (Noguchi et al.,
4.2.2. Photochemical recovery of NAD(P) from [NAD(P)]2 1997).
Bresnahan and Elving report that solutions containing the dimer Czochralska et al. reported that [NADP]2 oxidizes to NADP and
[NAD]2 are characteristically yellow-green in color. They observed that NADPH2 when irradiated at 365 nm with concomitant production of
upon exposure to daylight, the dimer solutions lost their yellow-green H2O2. They found that NADPH2 is also subject to oxidation by light, and
color in a few minutes while the solution in their dark electrolytic that its photooxidation is strongly dependent on the presence of oxygen
chamber remained yellow-green. They posited that a photochemical whereas the rate of photooxidation of [NADP]2 is similar under aerobic
reaction was responsible for the loss of dimer. They report that they and anaerobic conditions. The fact that the rate of anaerobic [NADP]2
illuminated their dimer solutions with UV light at 254 nm or at wave- photooxidation is competitive with aerobic [NADP]2 photooxidation
lengths exceeding 320 nm and found that the illuminated dimer solu- challenges the aerobic mechanism proposed by Liberatore et al. for
tion produced NAD in stoichiometric amounts. They also suggested that [NAD]2 photooxidation. Based on their experimental work, Czochralska
the reaction would produce some isomer NADH2, but their results did et al. proposed the following reaction scheme for the anaerobic pho-
not corroborate this hypothesis (Bresnahan and Elving, 1981). toconversion of [NADP]2 at 365 nm as shown in Fig. 8a and b.
Avigliano et al. described the light-induced homolytic cleavage of While the authors state that it is unclear what intermediates would
[NAD]2 dimer to NAD% free radicals which are further oxidized by be responsible for NADP and H2O2 production, they maintain that it
oxygen to NAD. They reported significant concomitant oxygen con- must be due to some interaction with water (Czochralska et al., 1990).
sumption and H2O2 production when [NAD]2 was exposed to UV light Photochemical recovery of NAD(P) from [NAD(P)]2 could be simi-
above 300 nm. These observations led them to propose the following larly engineered into an electrochemical bioreactor system to achieve
overall reaction: the same advantages as an enzymatic or electrochemical recovery re-
action as mentioned in Sections 4.2.3 and 4.2.4.
[NAD]2 + O2 + 2H+ → 2NAD + H2 O2 Reaction (3)

and the reaction products were confirmed with enzymatic assays. The 4.2.3. Electrochemical recovery of NAD(P) from [NAD(P)]2
proposed mechanism is as shown in Fig. 7. Dimeric species can be electrochemically oxidized to NAD(P).
[NAD]2 is put into an excited state by photon absorption. The ex- Specifically, Carelli et al. reported that the dimers undergo oxidation at
cited dimer then breaks up into NAD% free radicals, which can either − 0.2 V vs. SCE on a hanging mercury drop electrode, 0.5 V vs. SCE on
interact with oxygen to form the oxidized cofactor NAD and H2O2, or a platinum electrode, and 0.6 V vs. SCE on a glassy carbon electrode
they can interact with themselves to reform the dimer. Indeed, anae- (Carelli et al., 1981). Jaegfeldt reported that the dimers can be elec-
robic pathways for [NAD]2 photooxidation involving the formation of trochemically oxidized to NAD(P) at − 0.1 V vs. SCE on a mercury
dihydopyridyl radicals via interaction with solvent have been proposed, electrode. The author also adds that following electrochemical oxida-
but it is not obvious what intermediate would behave as an electron tion, the product was tested for activity against alcohol dehydrogenase
and ethanol. With this assay, 100% of the resulting NAD was en-
h zymatically active (Jaegfeldt, 1981). Similar results for the oxidation of
[NAD]2 [NAD]2* [NAD]2 and [NADP]2 dimer were reported by Schmakel et al. (1975).

4.2.4. Enzymatic recovery of NAD(P) from [NAD(P)]2


[NAD]2* 2 NAD Kôno and Suekane reported in 1958 that crude extracts from etio-
lated mung bean seedlings seemingly (but incorrectly) catalyzed the
complete oxidation of all NADH2 isomers to NAD. This was significant
2 NAD + O2 NAD + O2 - given that prior reports had determined that alcohol dehydrogenase,
malic acid dehydrogenase, and lactic acid dehydrogenase could only
partially oxidize the isomers. The discovery of a new enzyme that
NAD + NAD [NAD]2 completed the reaction was promising, but the enzyme they identified
required an unknown cofactor and was stimulated by methylene blue.
Based on their observations, they categorized it as a diaphorase despite
2 O2 - + 2 H+ H2O2 + O2 their experiments involving yeast or porcine diaphorases that showed
that only the mung bean enzyme can completely oxidize the substrate
Fig. 7. The proposed mechanism for the light-induced homolytic cleavage of [NAD]2, (Kôno and Suekane, 1958).
adapted from (Avigliano et al., 1983). In this scheme, [NAD]2 is oxidized to NAD in the Burnett and Underwood (1968) expanded on the work by Kôno and
presence of oxygen.
Suekane, and corrected Kôno's assumption that the mung bean enzyme

8
C.S. Morrison et al. Biotechnology Advances xxx (xxxx) xxx–xxx

Fig. 8. a. The proposed overall reactions for the light-in-


a duced anaerobic cleavage of [NADP]2, adapted from
[NADP]2 + H2O NADPH2 + NADP + OH (Czochralska et al., 1990). In this scheme, [NADP]2 un-
dergoes homolytic cleavage and rearrangement to form
NADP and NADPH2. b. The proposed mechanism for the
light-induced anaerobic cleavage of [NADP]2, adapted
[NADP]2 + H2O NADP + NADPH2 + OH- from (Czochralska et al., 1990). In this scheme, a photo-
excited [NADP]2 can react to form NADP and NADPH2.

b
h
[NADP]2 [NADP]2*

[NADP]2* 2 NADP

NADP + NADP [NADP]2

NADP + H2O NADPH2 + OH

h
2 [NADP]2 2 [NADP]2*

2 [NADP]2* + H+ NADP + further photochemical reactions

used NADH2 isomers as its substrate. They demonstrated experimen- from [NAD(P)]2 could be engineered into an electrochemical bioreactor
tally that the mung bean enzyme actually catalyzed the oxidation of system to deal with the adventitious formation of dimer that would
[NAD]2 to NAD. Fricks et al. reported that the catalytic activity of the otherwise lock all input NAD(P) cofactor material into an unusable
enzyme was greatly accelerated upon addition of phenol as Kôno's form and severely limit the number of turnovers theoretically achiev-
unknown cofactor to the reaction mixture, and thus they reclassified the able by the 1,4-NAD(P)H2 regeneration system (Morrison et al., 2017).
enzyme as a phenol oxidase. Further, they found analogs of the enzyme
in various other plants such as corn, cotton, wheat, and other beans.
However, the activity of the enzyme from crude extracts was lost upon 4.2.5. Other recovery reactions
purification. They reported that commercially-available mushroom Molecular oxygen is known to oxidize [NAD(P)]2 to NAD(P) with
polyphenol oxidase also catalyzed the oxidation of [NAD]2 to NAD increasing oxidation rates at higher temperatures and lower pH. In an
(Fricks et al., 1973). anoxic environment, [NAD(P)]2 has been reportedly stable for several
Carelli et al. reported that cytoplasmic fractions from rat liver cells days (Kovář et al., 1985). This knowledge was gained in the context of
stimulated the uptake of oxygen by [NAD]2, and they verified the [NAD(P)]2 stability in various storage conditions, so it is unlikely that
production of NAD and H2O2 by an enzyme assay (Carelli et al., 1980). molecular oxygen will be a good choice of method for intentional NAD
Avigliano et al. showed that horseradish peroxidase (HRP) catalyzes (P) recovery from [NAD(P)]2 since the rate of [NAD(P)]2 oxidation is on
the oxidation of [NAD]2 to NAD in a pH-dependent manner, which the time scale of several days.
suggests a free radical mechanism similar to that known for the oxi- Doxorubicin, an anthracycline chemotherapeutic drug, has been
dation and peroxidation of NADH2 to NAD under the same conditions reported to oxidize [NAD]2 to NAD in anaerobic conditions when ex-
(Avigliano et al., 1985). Two reaction schemes were proposed – one posed to light. It is thought to occur through a series of one-electron
consuming molecular oxygen as the oxidant, the other using hydrogen transfers involving NAD% radicals. The same reaction is not observed
peroxide. In either case, the result is conversion of the [NAD]2 dimer to when 1,4-NADH2 is used in place of [NAD]2 (Carelli et al., 1988). Given
two molecules of NAD plus a water molecule. Interestingly, bacterial its clinical use and deleterious health effects in humans, it is unlikely
NADH2 peroxidase was unable to catalyze the oxidation of [NAD]2 to that doxorubicin will be a good choice for NAD(P) recovery from [NAD
NAD. The authors suggest it is because bacterial NADH2 peroxidase (P)]2 in larger-scale applications.
involves a two-electron transfer, but [NAD]2 oxidation is only catalyzed One-electron-accepting proteins, such as cytochrome c, azurin and
by one-electron transfers which further supports the hypothesis that the methemoglobin, have also reportedly been able to oxidize [NAD(P)]2 to
responsible mechanism involves free radicals. Kirkor and Scheeline also NAD(P) via a reaction involving NAD(P)% radicals. Avigliano et al. re-
reported on the enzymatic activity of HRP with [NAD]2 to form NAD. In port that while [NAD(P)]2 reacts generally favorably with one-electron-
addition, they reported that [NAD]2 participates in other biological acceptors, the dimers are unable to react with two-electron-acceptors.
redox reactions that allows for the reintroduction of HRP into the os- While these proteins may not be the best choices for engineering NAD
cillatory peroxidase-oxidase cycle (Kirkor and Scheeline, 2000). (P) recovery reactions in larger-scale applications, the dimers may in-
While these reports have hinted at the potential biological sig- stead be suitable species for discriminating between one-electron-ac-
nificance of [NAD(P)]2 in primary metabolism, none to date have ceptors and two-electron-acceptors in biological redox reactions
identified the potential implications of enzymatic [NAD(P)]2 recovery (Avigliano et al., 1986).
methods to industrial applications. Similar to the case for renalase, Czochralska et al. reported that [NADP]2 is also oxidized to NADP
immobilized enzymes that are able to catalyze the recovery of NAD(P) by iron chelates. Fe(CN)63 − oxidized [NADP]2 in the dark due simply
to the more positive redox potential of the Fe(CN)63 −/Fe(CN)64 −

9
C.S. Morrison et al. Biotechnology Advances xxx (xxxx) xxx–xxx

couple than that of the NAD(P)/NAD(P)% couple. They also found that notion that the redox states of cells in an actively fermenting culture
Fe(EDTA)2 + and Fe(CN)63 − were able to increase the rate of [NADP]2 can be fundamentally modified in order to drive up the production of a
oxidation to NADP upon irradiation at 365 nm (Czochralska et al., reduced fermentation product. This is thought to occur by reducing the
1990). It follows, then, that other redox pairs with redox potentials amount of carbonaceous feedstock that is sacrificially oxidized by the
more positive than that of the NAD(P)/NAD(P)% couple should be able cells to provide electrons for the reduction of NAD(P) to 1,4-NAD(P)H2,
to chemically oxidize [NAD(P)]2 to NAD(P). While this may be inter- thereby increasing the efficiency of carbon flux through metabolic
esting to some researchers hoping to recover NAD(P) from NAD(P) in pathways toward the target product. For example, the yield of 1,3-
small-scale applications, it is unlikely that this strategy could be em- propanediol by a mixed culture growing on glycerol was enhanced from
ployed for larger scale cofactor regeneration since the addition of extra 24.8% to 50.1% when an external current was applied to the culture.
chemicals to the system constitutes an added material cost in addition The authors found through a metabolic flux analysis that glycerol me-
to adding to the cost and complexity of downstream purifications. tabolism was redirected from propionate fermentation to 1,3-propane-
In summary, it is now well established in the literature that various diol production (Zhou et al., 2013). An example of a process catalyzed
enzymatic, chemical, and photochemical dimer oxidation mechanisms by a recombinant culture is the production of biofuels by Ralstonia
involve the formation of NAD(P)% radicals in aerobic and anaerobic eutropha. The authors reported the electrically-driven generation of
reactions. This may point to some fundamental biochemical property of 140 mg L− 1 of biofuels from CO2 alone (Li et al., 2012).
the dimer that is important to redox processes involving NAD(P)H2 However, the mechanisms by which electrically-driven cultures in-
(Avigliano et al., 1985; Avigliano et al., 1983; Czochralska et al., 1990; crease the carbon efficiencies and yields of reduced fermentation pro-
Carelli et al., 1988; Avigliano et al., 1986). ducts is not well understood. Harrington et al. looked at the metabolite
profiles of Escherichia coli, Klebsiella pneumoniae, and Zymomonas mobilis
following neutral red-mediated electrosynthesis to observe changes in
5. Conclusions
the organisms' metabolisms during electrically-driven fermentations.
They found through stoichiometric analysis that there were significant
5.1. Opportunities for in vitro and in vivo biocatalysis
metabolite changes observed during electrosynthesis for E. coli and K.
pneumoniae. No changes in Z. mobilis metabolism were observed. While
5.1.1. P450 monooxygenase reactions
the authors claim that NAD reduction alone could not account for all
There are many other potential applications beyond ketoreductases.
the metabolite changes observed, it is clear from their work that a
The P450 monooxygenases, which require a reducing equivalent to take
variety of electrically-driven effects on microbial metabolism culminate
up one oxygen atom from an oxygen molecule transforming it to water,
in an increase of carbon efficiency and yield of fermentation products
are very important reactions in living systems that are responsible for
(Harrington et al., 2015). There is a significant need for more research
inserting an oxygen atom between a carbon hydrogen bond (Urlacher
to be done to understand the mechanisms by which some organisms are
and Eiben, 2006; Vilker et al., 1999; Leonard and Koffas, 2007; Leonard
able to interact with their external environment to alter their own in-
et al., 2006; Leonard et al., 2005). There are no known chemical cat-
tracellular redox environment. A clearer fundamental understanding of
alysts that are capable of accomplishing this transformation that is es-
these phenomena could lead to the development of industrial microbial
sential to steroid hydroxylation, metabolite formation, and general
biological degradation of organic materials in the environment (see strains that are engineered to optimize target production without in-
creasing the amount of input carbonaceous material in electrically-
Fig. 9).
driven fermentations.
Although more work remains to be done to understand microbial
5.1.2. Microbial electrosynthesis electrosynthesis before extensive engineering can take place, the reader
The application of electrochemical bioreactors to the field of mi- is encouraged to read the review by White et al. for a current, com-
crobial whole-cell electrosynthesis is not comprehensively reviewed in prehensive understanding of how bacterial extracellular electron ex-
this paper. However, it deserves to be mentioned in order to make the change is thought to occur (White et al., 2016).
reader aware of the applications of electrochemical bioreactors in the
context of metabolic engineering, synthetic biology, and industrial
microbiology. The reader who is interested in microbial electrosynth- 5.1.3. In vitro biocatalysis
esis is encouraged to read some recent review articles that have given For the first time, strategies for dealing with the adventitious for-
an extensive treatment of the topic (Kato, 2015; Mohan et al., 2014; mation of useless and inhibitory reduced cofactor side products are
Choi and Sang, 2016; TerAvest and Ajo-Franklin, 2016; Tremblay and presented in the context of applied biochemistry and industrial bioca-
Zhang, 2015; Harnisch et al., 2015). talysis. Such reactions are henceforth designated as “recovery reac-
In general, the concept of microbial electrosynthesis is driven by the tions” and are crucial to the success of engineered electrochemical NAD
(P)H2 regeneration systems. Recovery reactions should be used in
R1 tandem with operational controls to minimize or eliminate the forma-
R1 H
H O tion of reduced cofactor side products while simultaneously maximizing
H H O2 H2O the selective formation of 1,4-NAD(P)H2.
R2
R2
R1
R1
P450 5.2. Current state-of-the-art for electrochemical reactor technology
O
enzyme
R2 To date, engineered electrochemical bioreactor systems have in-
R2
evitably suffered from the nonspecific reduction problem described in
R1 Section 2.2. It has not been possible to engineer an efficient or an
R1 O
O 1,4-NAD(P)H2 NAD(P) economic system for the concomitant in situ regeneration of 1,4-NAD
O
R2 (P)H2 with the formation of a product of interest via the catalytic ac-
R2
tivity of a redox enzyme. This review therefore serves as a path toward
Fig. 9. The utility of 1,4-NAD(P)H2-requiring P450 enzymes for biocatalysis is high- engineered solutions by which to achieve efficient and economic large-
lighted through various chemical transformations. The cost-efficient regeneration of 1,4- scale 1,4-NAD(P)H2-dependent biocatalytic processes using electro-
NADPH2 is essential to give access to these reactions on a meaningful scale.
chemical methods.

10
C.S. Morrison et al. Biotechnology Advances xxx (xxxx) xxx–xxx

5.3. Future outlook Acknowledgments

Many practical challenges remain for the electrochemical re- This work was supported by BioChemInsights, Inc., Malvern, PA
generation of 1,4-NAD(P)H2, and this paper concludes by calculating through a Sponsored Research Agreement with Rensselaer Polytechnic
some benchmarks which should be kept in mind for practical systems. Institute; and the National Science Foundation Graduate Research
Generating hydrogen gas by conventional electrolysis using power Fellowship Program [grant number DGE-1247271].
from the electrical grid at US$60/KW gives a cost approximately US
$3200–$3500/mt. If hydrogen is supplied within a biological process References
by the metabolism of glucose having a cost of US$450/mt, and running
a “costless” process by which glucose is metabolized completely to CO2 Ali, I., Gill, A., Omanovic, S., 2012a. Direct electrochemical regeneration of the enzymatic
cofactor 1,4-NADH employing nano-patterned glassy carbon/Pt and glassy carbon/Ni
and 1,4-NAD(P)H2, the equivalent hydrogen cost is US$6750/mt.
electrodes. Chem. Eng. J. 188, 173–180. http://dx.doi.org/10.1016/j.cej.2012.02.
If 1,4-NAD(P)H2 is generated by oxidizing glucose to glucono- 005.
lactone in a coupled enzyme system using glucose dehydrogenase, the Ali, I., McArthur, M., Hordy, N., Coulombe, S., Omanovic, S., 2012b. Electrochemical
regeneration of the cofactor NADH employing a carbon nanofibers cathode. Int. J.
equivalent hydrogen cost is US$40,500/mt. Replacing glucose with Electrochem. Sci. 7, 7675–7683.
sodium formate at a cost of US$250/mt and using formate dehy- Ali, I., Soomro, B., Omanovic, S., 2011. Electrochemical regeneration of NADH on a glassy
drogenase to generate 1,4-NAD(P)H2 and CO2, the equivalent hydrogen carbon electrode surface: the influence of electrolysis potential. Electrochem.
Commun. 13, 562–565. http://dx.doi.org/10.1016/j.elecom.2011.03.010.
cost is US$8500/mt. Using ethanol at the current price of US$1.50/ Angelastro, A., Dawson, W.M., Luk, L.Y.P., Allemann, R.K., 2017. A versatile disulfide-
USgal, or US$500/mt, in a process in which acetaldehyde is generated, driven recycling system for NADP+ with high cofactor turnover number. ACS Catal.
7, 1025–1029. http://dx.doi.org/10.1021/acscatal.6b03061.
the equivalent hydrogen cost is US$16,000/mt. This could be reduced Avigliano, L., Carelli, V., Casini, A., Finazzi-Agrò, A., Liberatore, F., 1983. Photocatalyzed
to US$8000/mt hydrogen if the ethanol were oxidized to acetic acid. oxidation of NAD dimers. Biochim. Biophys. Acta Bioenerg. 723, 372–375. http://dx.
Using an electrochemical bioreactor with the same efficiency as a doi.org/10.1016/0005-2728(83)90043-9.
Avigliano, L., Carelli, V., Casini, A., Finazzi-Agrò, A., Liberatore, F., 1985. Oxidation of
current electrolyzer would give the same equivalent hydrogen cost; US NAD dimers by horseradish peroxidase. Biochem. J. 226, 391–395.
$3200–3500/mt, but this would be delivered as 1,4-NAD(P)H2 to the Avigliano, L., Carelli, V., Casini, A., Finazzi-Agrò, A., Liberatore, F., Rossi, A., 1986.
enzyme of interest. Like a standard electrolyzer, this allows the current Oxidation of nicotinamide coenzyme dimers by one-electron-accepting proteins.
Biochem. J. 237.
electrical grid to be used as a fungible hydrogen source, with any Azem, A., Man, F., Omanovic, S., 2004. Direct regeneration of NADH on a ruthenium
proportion of power from renewable sources that may be available. modified glassy carbon electrode. J. Mol. Catal. A Chem. 219, 283–299. http://dx.
doi.org/10.1016/j.molcata.2004.04.041.
Of course, the cofactor material itself has a cost, and at present this Baik, S.H., Kang, C., Jeon, I.C., Yun, S.E., 1999. Direct electrochemical regeneration of
is approximately US$1200/kg for NAD. One mole of NAD can be loaded NADH from NAD + using cholesterol-modified gold amalgam electrode. Biotechnol.
with one mole of hydrogen as 1,4-NADH2, and with a molecular weight Tech. 13, 1–5. http://dx.doi.org/10.1023/A:1008865212773.
Beaupre, B.A., Carmichael, B.R., Hoag, M.R., Shah, D.D., Moran, G.R., 2013. Renalase is
of 665 for 1,4-NADH2, that would give a cofactor cost of just under US an α-NAD(P)H oxidase/anomerase. J. Am. Chem. Soc. 135, 13980–13987. http://dx.
$500 per mol of reduced cofactor – completely swamping any con- doi.org/10.1021/ja407384h.
Beaupre, B.A., Hoag, M.R., Roman, J., Försterling, F.H., Moran, G.R., 2015. Metabolic
sideration of the cost of the hydrogen that mol of cofactor is carrying!
function for human renalase: oxidation of isomeric forms of β-NAD(P)H that are
Hence the importance of recycling the cofactor material, and ensuring it inhibitory to primary metabolism. Biochemistry 54, 795–806. http://dx.doi.org/10.
is all recovered as NAD ready to be put through the reduction cycle 1021/bi5013436.
Biellmann, J.F., Lapinte, C., Haid, E., Weimann, G., 1979. Structure of lactate dehy-
again. This will be true for any in vitro system that is used to perform drogenase inhibitor generated from coenzyme. Biochemistry 18, 1212–1217. http://
cofactor recycling, enzymatic or non-enzymatic. With 1000 recycles of dx.doi.org/10.1021/bi00574a015.
the cofactor, the cost contribution of the cofactor material drops to US Bresnahan, W., Elving, P., 1981. Spectrophotometric investigation of products formed
following the initial one-electron electrochemical reduction of nicotinamide adenine
$0.50 per mol of delivered hydrogen. But with 500,000 mols of hy- dinucleotide (NAD+). Biochim. Biophys. Acta Gen. Subj. 678, 151–156. http://dx.
drogen per mt, this is still US$250,000 cofactor cost per mt of delivered doi.org/10.1016/0304-4165(81)90200-2.
Burnett, R.W., Underwood, A.L., 1968. A dimer of diphosphopyridine nucleotide.
hydrogen. Biochemistry 7, 3328–3333. http://dx.doi.org/10.1021/bi00850a003.
Products in which one mole of hydrogen is consumed in a reaction Carelli, V., Casini, A., Finazzi-Agro, A., Liberatore, F., Tortorella, S., 1988. Photocatalyzed
giving a product with a molecular weight of 400, and with a cofactor anaerobic oxidation of nicotinamide coenzyme dimers to NAD+ by adriamycin. Free
Radic. Res. Commun. 4, 397–402.
recycle of 1000 times, gives a cofactor cost of contribution of just over Carelli, V., Liberatore, F., Casini, A., Mondelli, R., Arnone, A., Carelli, I., Rotilio, G.,
US$100 per kilogram of product. This is well within the acceptable cost Mavelli, I., 1980. Dimers of nicotinamide adenine dinucleotide: new evidence for the
range of a process step in the synthesis of a pharmaceutical, and at structure and the involvement in an enzymatic redox process. Bioorg. Chem. 9,
342–351. http://dx.doi.org/10.1016/0045-2068(80)90044-9.
10,000 turnovers of the cofactor, would give a step cost acceptable for Carelli, I., Rosati, R., Casini, A., 1981. Electrochemical oxidation of dimers of nicotina-
use in high value specialty products. mide adenine dinucleotide in aqueous medium. Electrochim. Acta Pergamon. http://
dx.doi.org/10.1016/0013-4686(81)85149-3.
It is anticipated that laboratory-scale in vitro electrochemical bior- Chenault, H.K., Simon, E.S., Whitesides, G.M., 1988. Cofactor regeneration for enzyme-
eactor systems capable of routinely supplying g/hour stoichiometric catalysed synthesis. Biotechnol. Genet. Eng. Rev. 6, 221–270. http://dx.doi.org/10.
quantities of 1,4-NAD(P)H2 will become practical in the near future. 1080/02648725.1988.10647849.
Chenault, H.K., Whitesides, G.M., 1987. Regeneration of nicotinamide cofactors for use in
These systems will significantly lower the cost of 1,4-NAD(P)H2 while organic synthesis. Appl. Biochem. Biotechnol. 14, 147–197. http://dx.doi.org/10.
facilitating easy use of 1,4-NAD(P)H2 dependent enzymes in potentially 1007/BF02798431.
Choi, O., Sang, B.-I., 2016. Extracellular electron transfer from cathode to microbes:
useful enzyme reactions.
application for biofuel production. Biotechnol. Biofuels 9, 11. http://dx.doi.org/10.
The larger-scale commercial production of 1,4-NAD(P)H2-depen- 1186/s13068-016-0426-0.
dent products will be on the horizon with the first processes most like Czochralska, B., Bojarska, E., Pawlicki, K., Shugar, D., 1990. Photochemical and enzy-
matic redox transformations of reduced forms of coenzyme NADP +. Photochem.
involving P450 monooxygenase and other high value-added pharma- Photobiol. 51, 401–410. http://dx.doi.org/10.1111/j.1751-1097.1990.tb01731.x.
ceutical intermediates. Eikeren, P. Van, 1989. Use and regeneration of reagents using coupled reactions and
permselective barriers. EP0349204 A3.
Elving, P.J., Bresnahan, W.T., Moiroux, J., Samec, Z., 1982. NAD/NADH as a model redox
system: mechanism, mediation, modification by the environment. Bioelectrochem.
Conflict of interest statement Bioenerg. 9, 365–378. http://dx.doi.org/10.1016/0302-4598(82)80026-3.
Fricks, D.H., Bechtel, J.T., Underwood, A.L., 1973. Reactivity of nicotinamide-adenine
dinucleotide dimer in plant phenol oxidase systems. Arch. Biochem. Biophys. 159,
The authors declare that CSM, WBA, DRD, and MAGK are co-in- 837–841. http://dx.doi.org/10.1016/0003-9861(73)90525-0.
ventors on a patent application related to the application of renalase to Godtfredsen, S.E., Ottesen, M., 1978. 1,6-dihydro-NAD as an humidity-induced lactate
dehydrogenase inhibitor in NADH preparations. Carlsb. Res. Commun. 43, 171–175.
electrochemical bioreactor systems. http://dx.doi.org/10.1007/BF02914239.
Harnisch, F., Rosa, L.F.M., Kracke, F., Virdis, B., Krömer, J.O., 2015. Electrifying white
biotechnology: engineering and economic potential of electricity-driven bio-

11
C.S. Morrison et al. Biotechnology Advances xxx (xxxx) xxx–xxx

production. ChemSusChem 8, 758–766. http://dx.doi.org/10.1002/cssc.201402736. Morrison, C.S., Armiger, W.B., Dodds, D.R., Koffas, M.A.G., 2017. Improved Method for
Harrington, T.D., Mohamed, A., Tran, V.N., Biria, S., Gargouri, M., Park, J.-J., Gang, D.R., Using Electrochemical Bioreactor Module With Recovery of Cofactor.
Beyenal, H., 2015. Neutral red-mediated microbial electrosynthesis by Escherichia (WO2017160793).
coli, Klebsiella pneumoniae, and Zymomonas mobilis. Bioresour. Technol. 195, Noguchi, N., Tachikawa, M., Takahashi, H., 1997. One-electron photooxidation of Nadh:
57–65. http://dx.doi.org/10.1016/j.biortech.2015.06.005. laser photolysis, time-resolved raman and Ab initio Mo calculation study. In:
Hildebrand, F., Lütz, S., 2007. Electroenzymatic synthesis of chiral alcohols in an Spectroscopy of Biological Molecules: Modern Trends. Springer, Netherlands,
aqueous–organic two-phase system. Tetrahedron Asymmetry 18, 1187–1193. http:// Dordrecht, pp. 161–162. http://dx.doi.org/10.1007/978-94-011-5622-6_71.
dx.doi.org/10.1016/j.tetasy.2007.05.002. Pandini, V., Ciriello, F., Tedeschi, G., Rossoni, G., Zanetti, G., Aliverti, A., 2010. Synthesis
Hoag, M.R., Roman, J., Beaupre, B.A., Silvaggi, N.R., Moran, G.R., 2015. Bacterial re- of human renalase1 in Escherichia coli and its purification as a FAD-containing ho-
nalase: structure and kinetics of an enzyme with 2- and 6-dihydro-β-NAD(P) oxidase loprotein. Protein Expr. Purif. 72, 244–253. http://dx.doi.org/10.1016/j.pep.2010.
activity from Pseudomonas phaseolicola. Biochemistry 54, 3791–3802. http://dx.doi. 03.008.
org/10.1021/acs.biochem.5b00451. Patel, R., 2006. Biocatalysis: synthesis of chiral intermediates for pharmaceuticals. Curr.
Hollmann, F., Hofstetter, K., Schmid, A., 2006. Non-enzymatic regeneration of nicotina- Org. Chem. 10, 1289–1321. http://dx.doi.org/10.2174/138527206777698011.
mide and flavin cofactors for monooxygenase catalysis. Trends Biotechnol. 24, Rabaey, K., Angenent, L., Schroder, U., Keller, J., 2010. Bioelectrochemical Systems:
163–171. http://dx.doi.org/10.1016/j.tibtech.2006.02.003. From Extracellular Electron Transfer to Biotechnological Application, 1st ed. IWA
Hollmann, F., Schmid, A., Steckhan, E., 2001. The first synthetic application of a Publishing, London.
monooxygenase employing indirect electrochemical NADH regeneration. Angew. Schmakel, C.O., Santhanam, K.S.V., Elving, P.J., 1975. Nicotinamide adenine dinucleo-
Chem. Int. Ed. 40, 169–171. http://dx.doi.org/10.1002/1521-3773(20010105) tide (NAD +) and related compounds. Electrochemical redox pattern and allied
40:1<169::AID-ANIE169>3.0.CO;2-T. chemical behavior. J. Am. Chem. Soc. 97, 5083–5092. http://dx.doi.org/10.1021/
Homann, M.J., Previte, E., 1996. Stereoselective microbial reduction of 5-fluorophenyl-5- ja00851a010.
oxo-pentanoic acid and a phenyloxazolidinone condensation product thereof. Schroer, K., Lütz, S., 2009. A continuously operated bimembrane reactor process for the
US5618707 A. biocatalytic production of (2R,5R)-hexanediol. Org. Process. Res. Dev. 13,
Jaegfeldt, H., 1981. A study of the products formed in the electrochemical reduction of 1202–1205. http://dx.doi.org/10.1021/op9001643.
nicotinamide-adenine-dinucleotide. Bioelectrochem. Bioenerg. 8, 355–370. http:// Schroer, K., Mackfeld, U., Tan, I.A.W., Wandrey, C., Heuser, F., Bringer-Meyer, S.,
dx.doi.org/10.1016/0302-4598(81)80018-9. Weckbecker, A., Hummel, W., Daußmann, T., Pfaller, R., Liese, A., Lütz, S., 2007.
Kang, H.S., Na, B.K., Park, D.H., 2007. Oxidation of butane to butanol coupled to elec- Continuous asymmetric ketone reduction processes with recombinant Escherichia
trochemical redox reaction of NAD +/NADH. Biotechnol. Lett. 29, 1277–1280. Coli. J. Biotechnol. 132, 438–444. http://dx.doi.org/10.1016/j.jbiotec.2007.08.003.
http://dx.doi.org/10.1007/s10529-007-9385-7. Shin, I.H., Jeon, S.J., Park, H.S., Park, D.H., 2004. Catalytic oxidoreduction of pyruvate/
Kato, S., 2015. Biotechnological aspects of microbial extracellular electron transfer. lactate and acetaldehyde/ethanol coupled to electrochemical oxidoreduction of NAD
Microbes Environ. 30, 133–139. http://dx.doi.org/10.1264/jsme2.ME15028. +/NADH. J. Microbiol. Biotechnol. 14, 540–546.
Kirkor, E.S., Scheeline, A., 2000. Nicotinamide adenine dinucleotide species in the Siu, E., Won, K., Park, C.B., 2007. Electrochemical regeneration of NADH using con-
horseradish peroxidase-oxidase oscillator. Eur. J. Biochem. 267, 5014–5022. http:// ductive vanadia-silica xerogels. Biotechnol. Prog. 23, 293–296. http://dx.doi.org/10.
dx.doi.org/10.1046/j.1432-1327.2000.01554.x. 1021/bp060247l.
Kôno, T., Suekane, M., 1958. A diphosphopyridine nucleotide oxidase in mung bean Ströhle, F.W., Kranen, E., Schrader, J., Maas, R., Holtmann, D., 2016. A simplified process
seedlings. Bull. Agric. Chem. Soc. Jpn. 22 (6), 404–410. design for P450 driven hydroxylation based on surface displayed enzymes.
Kovář, J., Turánek, J., Hlaváč, C., Vala, V., Kahle, V., 1985. Liquid chromatographic se- Biotechnol. Bioeng. 113, 1225–1233. http://dx.doi.org/10.1002/bit.25885.
parations of dimers of nicotinamide adenine dinucleotide and nicotinamide adenine TerAvest, M.A., Ajo-Franklin, C.M., 2016. Transforming exoelectrogens for biotechnology
dinucleotide phosphate. J. Chromatogr. A 319, 341–349. http://dx.doi.org/10.1016/ using synthetic biology. Biotechnol. Bioeng. 113, 687–697. http://dx.doi.org/10.
S0021-9673(01)90570-9. 1002/bit.25723.
Lau, C., Flechsig, G.-U., Gründler, P., Wang, J., 2005. Electrochemistry of nicotinamide Tremblay, P.-L., Zhang, T., 2015. Electrifying microbes for the production of chemicals.
adenine dinucleotide (reduced) at heated platinum electrodes. Anal. Chim. Acta 554, Front. Microbiol. 6, 201. http://dx.doi.org/10.3389/fmicb.2015.00201.
74–78. http://dx.doi.org/10.1016/j.aca.2005.08.044. Ullah, N., Ali, I., Omanovic, S., 2015. Direct electrocatalytic reduction of coenzyme NAD
Lee, L.G., Whitesides, G.M., 1985. Enzyme-catalyzed organic synthesis: a comparison of + to enzymatically-active 1,4-NADH employing an iridium/ruthenium-oxide elec-
strategies for in situ regeneration of NAD from NADH. J. Am. Chem. Soc. 107, trode. Mater. Chem. Phys. 149–150, 413–417. http://dx.doi.org/10.1016/j.
6999–7008. http://dx.doi.org/10.1021/ja00310a043. matchemphys.2014.10.038.
Leonard, E., Koffas, M.A.G., 2007. Engineering of artifical plant cytochrome P450 en- Urlacher, V.B., Eiben, S., 2006. Cytochrome P450 monooxygenases: perspectives for
zymes for synthesis of isoflavones by Escherichia coli. Appl. Environ. Microbiol. 73, synthetic application. Trends Biotechnol. 24, 324–330. http://dx.doi.org/10.1016/j.
7246–7251. tibtech.2006.05.002.
Leonard, E., Yan, Y., Koffas, M.A., 2006. Functional expression of a P450 flavonoid hy- Vilker, V.L., Reipa, V., Mayhew, M., Holden, M.J., 1999. Challenges in capturing oxy-
droxylase for the biosynthesis of plant-specific hydroxylated flavonols in Escherichia genase activity in vitro. J. Am. Oil Chem. Soc. 76, 1283–1289. http://dx.doi.org/10.
coli. Metab. Eng. 8, 172–181. 1007/s11746-999-0140-1.
Leonard, E., Yan, Y., Lim, K.H., Koffas, M.A., 2005. Investigation of two distinct flavone Walsh, F., Reade, G., 1994. Design and performance of electrochemical reactors for ef-
synthases for plant-specific flavone biosynthesis in Saccharomyces cerevisiae. Appl. ficient synthesis and environmental treatment. Part 2. Typical reactors and their
Environ. Microbiol. 71, 8241–8248. performance. Analyst 119, 797. http://dx.doi.org/10.1039/an9941900797.
Leuchtenberger, W., Huthmacher, K., Drauz, K., 2005. Biotechnological production of Wang, X., Yiu, H.H.P., 2016. Heterogeneous catalysis mediated cofactor NADH re-
amino acids and derivatives: current status and prospects. Appl. Microbiol. generation for enzymatic reduction. ACS Catal. 6, 1880–1886. http://dx.doi.org/10.
Biotechnol. 69, 1–8. http://dx.doi.org/10.1007/s00253-005-0155-y. 1021/acscatal.5b02820.
Li, H., Opgenorth, P.H., Wernick, D.G., Rogers, S., Wu, T.-Y., Higashide, W., Malati, P., White, G.F., Edwards, M.J., Gomez-Perez, L., Richardson, D.J., Butt, J.N., Clarke, T.A.,
Huo, Y.-X., Cho, K.M., Liao, J.C., 2012. Integrated electromicrobial conversion of CO2 2016. Mechanisms of bacterial extracellular electron exchange. Adv. Microb. Physiol.
to higher alcohols. Science 80 (335), 1596. http://dx.doi.org/10.1126/science. 68, 87–138. http://dx.doi.org/10.1016/bs.ampbs.2016.02.002.
1217643. Williams, J.N., 1952. Inhibition of coenzyme I-requiring enzyme by adenine and Adenyl
Long, Y.-T., Chen, H.-Y., 1997. Electrochemical regeneration of coenzyme NADH on a metabolites in vitro. J. Biol. Chem. 195, 629.
histidine modified silver electrode. J. Electroanal. Chem. 440, 239–242. http://dx. Wong, C.H., Whitesides, G.M., 1982. Enzyme-catalyzed organic synthesis: NAD(P)H co-
doi.org/10.1016/S0022-0728(97)80061-7. factor regeneration using ethanol/alcohol dehydrogenase/aldehyde dehydrogenase
Ma, S.M., Garcia, D.E., Redding-Johanson, A.M., Friedland, G.D., Chan, R., Batth, T.S., and methanol/alcohol dehydrogenase/aldehyde dehydrogenase/formate dehy-
Haliburton, J.R., Chivian, D., Keasling, J.D., Petzold, C.J., Soon Lee, T., Chhabra, drogenase. J. Org. Chem. 47, 2816–2818. http://dx.doi.org/10.1021/jo00135a037.
S.R., 2011. Optimization of a heterologous mevalonate pathway through the use of Wong, C.H., Whitesides, G.M., 1983. Enzyme-catalyzed organic synthesis: regeneration of
variant HMG-CoA reductases. Metab. Eng. 13, 588–597. http://dx.doi.org/10.1016/ deuterated nicotinamide cofactors for use in large-scale enzymatic synthesis of
j.ymben.2011.07.001. deuterated substances. J. Am. Chem. Soc. 105, 5012–5014. http://dx.doi.org/10.
Ma, S.K., Gruber, J., Davis, C., Newman, L., Gray, D., Wang, A., Grate, J., Huisman, G.W., 1021/ja00353a026.
Sheldon, R.A., 2010. A green-by-design biocatalytic process for atorvastatin inter- Xu, J., Li, G., Wang, P., Velazquez, H., Yao, X., Li, Y., Wu, Y., Peixoto, A., Crowley, S.,
mediate. Green Chem. 12, 81–86. http://dx.doi.org/10.1039/B919115C. Desir, G.V., 2005. Renalase is a novel, soluble monoamine oxidase that regulates
Mädje, K., Schmölzer, K., Nidetzky, B., Kratzer, R., 2012. Host cell and expression en- cardiac function and blood pressure. J. Clin. Investig. 115, 1275–1280.
gineering for development of an E. coli ketoreductase catalyst: enhancement of for- Zaks, A., Dodds, D.R., 1997. Application of biocatalysis and biotransformations to the
mate dehydrogenase activity for regeneration of NADH. Microb. Cell Fact 11, 7. synthesis of pharmaceuticals. Drug Discov. Today 2, 513–531. http://dx.doi.org/10.
http://dx.doi.org/10.1021/bp0200954. 1016/S1359-6446(97)01078-7.
Man, F., Omanovic, S., 2004. A kinetic study of NAD + reduction on a ruthenium mod- Zambianchi, F., Pasta, P., Carrea, G., Colonna, S., Gaggero, N., Woodley, J.M., 2002. Use
ified glassy carbon electrode. J. Electroanal. Chem. 568, 301–313. http://dx.doi.org/ of isolated cyclohexanone monooxygenase from recombinantEscherichia coli as a
10.1016/j.jelechem.2004.02.006. biocatalyst for Baeyer-Villiger and sulfide oxidations. Biotechnol. Bioeng. 78,
McClay, K., Fox, B.G., Steffan, R.J., 2000. Toluene monooxygenase-catalyzed epoxidation 489–496. http://dx.doi.org/10.1002/bit.10207.
of alkenes. Appl. Environ. Microbiol. 66, 1877–1882. http://dx.doi.org/10.1128/ Zhao, H., van der Donk, W.A., 2003. Regeneration of cofactors for use in biocatalysis.
AEM.66.5.1877-1882.2000. Curr. Opin. Biotechnol. 14, 583–589. http://dx.doi.org/10.1016/j.copbio.2003.09.
Mohan, S.V., Velvizhi, G., Vamshi Krishna, K., Lenin Babu, M., 2014. Microbial catalyzed 007.
electrochemical systems: a bio-factory with multi-facet applications. Bioresour. Zhou, M., Chen, J., Freguia, S., Rabaey, K., Keller, J., 2013. Carbon and electron fluxes
Technol. 165, 355–364. http://dx.doi.org/10.1016/j.biortech.2014.03.048. during the electricity driven 1,3-propanediol biosynthesis from glycerol. Environ. Sci.
Moran, G.R., 2016. The catalytic function of renalase: a decade of phantoms. Biochim. Technol. 11199–11205.
Biophys. Acta 1864, 177–186. http://dx.doi.org/10.1016/j.bbapap.2015.04.010.

12

You might also like