Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Electrochemical transformations catalyzed by cytochrome P450s and peroxidases

Neeraj Kumar a, Jie He *ab and James F. Rusling *abcd
aDepartment of Chemistry, University of Connecticut, Storrs, CT 06269-3136, USA. E-mail: jie.he@uconn.edu; james.rusling@uconn.edu
bInstitute of Materials Science, University of Connecticut, Storrs, CT 06269-3136, USA
cDepartment of Surgery and Neag Cancer Center, Uconn Health, Farmington, CT 06030, USA
dSchool of Chemistry, National University of Ireland at Galway, Galway, Ireland

Received 12th June 2023

First published on 17th July 2023


Abstract

Cytochrome P450s (Cyt P450s) and peroxidases are enzymes featuring iron heme cofactors that have wide applicability as biocatalysts in chemical syntheses. Cyt P450s are a family of monooxygenases that oxidize fatty acids, steroids, and xenobiotics, synthesize hormones, and convert drugs and other chemicals to metabolites. Peroxidases are involved in breaking down hydrogen peroxide and can oxidize organic compounds during this process. Both heme-containing enzymes utilize active FeIV[double bond, length as m-dash]O intermediates to oxidize reactants. By incorporating these enzymes in stable thin films on electrodes, Cyt P450s and peroxidases can accept electrons from an electrode, albeit by different mechanisms, and catalyze organic transformations in a feasible and cost-effective way. This is an advantageous approach, often called bioelectrocatalysis, compared to their biological pathways in solution that require expensive biochemical reductants such as NADPH or additional enzymes to recycle NADPH for Cyt P450s. Bioelectrocatalysis also serves as an ex situ platform to investigate metabolism of drugs and bio-relevant chemicals. In this paper we review biocatalytic electrochemical reactions using Cyt P450s including C–H activation, S-oxidation, epoxidation, N-hydroxylation, and oxidative N-, and O-dealkylation; as well as reactions catalyzed by peroxidases including synthetically important oxidations of organic compounds. Design aspects of these bioelectrocatalytic reactions are presented and discussed, including enzyme film formation on electrodes, temperature, pH, solvents, and activation of the enzymes. Finally, we discuss challenges and future perspective of these two important bioelectrocatalytic systems.


image file: d3cs00461a-p1.tif

Neeraj Kumar

Neeraj Kumar received his PhD in Chemistry from Indian Institute of Technology Roorkee, India (2019). Then he worked as Research Associate at the CSIR-Central Building Research Institute Roorkee and National Postdoctoral Fellow at the Department of Inorganic and Physical Chemistry, at the Indian Institute of Science (IISc) Bangalore. Currently, he works as a Postdoctoral Research Associate at the Department of Chemistry, University of Connecticut, Storrs, USA. His research interests include electrochemistry, electrochemical sensors/biosensor, photo-catalysis, and electro-/bio-catalysis.

image file: d3cs00461a-p2.tif

Jie He

Jie He received his BS and MS degrees in Polymer Materials Science and Engineering from Sichuan University and his PhD in Chemistry from Université de Sherbrooke in 2010. He joined the faculty of the University of Connecticut after working as a postdoctoral fellow at the University of Maryland in 2011–2013. He is currently an Associate Professor of Chemistry and Polymer Program at the Institute of Materials Science, University of Connecticut. His research group focuses on design and synthesis of hybrid materials of polymers and metals that are capable activating small molecules as inspired by metalloenzymes.

image file: d3cs00461a-p3.tif

James F. Rusling

James F. Rusling earned a BSc from Drexel University (1969), and PhD in Chemistry from Clarkson University (1979). He is Paul Krenicki Professor of Chemistry at University of Connecticut, Professor of Surgery and Neag Cancer Center member at UConn Health, and adjunct Professor of Chemistry at National Univ. of Ireland, Galway. Current research includes biocatalysis, electrochemical biocatalysis, microfluidic devices for point-of-care disease diagnostics, nanoenergy for implantable medical devices, and mass spectrometry methods for peptide biomarkers, He has authored over 450 research papers and several books, and is a musician interested in Irish and American folk styles.


1. Introduction

Chemical transformations of organic compounds are fundamentally and practically important for synthesizing essential chemical products like new and safe drugs, clean industrial chemicals, environmentally safe materials, and beyond.1 A key to making these transformations practical is high yield at low cost. In comparison to traditional chemical syntheses, biological syntheses are most often catalyzed by enzymes and very often outperform chemical catalysis in terms of substrate specificity and product regio- and stereo-selectivity. Thus, the utilization of biological catalysts in chemical transformations can provide cost-effective, selective, rapid syntheses, often without separation of by-products or intermediates.2,3

Novel synthetic methods are needed to make strictly chiral pharmaceutical products. For example, often only one chiral form of a drug has the desired activity, while the other chiral form may have undesirable or even toxic activity. Biocatalysts offer alternative routes to chemical syntheses under mild conditions, low environmental toxicity, and excellent chiral selectivity, and can avoid toxic solvents and redox agents. Enzymes lead to high regio- and stereoselectivity in the reaction products.1,3–5 The study of drug metabolites also plays a key role in uncovering therapeutic and toxicity effects.6,7 Heme-containing enzymes especially Cytochrome P450 (Cyt P450s, Scheme 1), are involved in up to 75% of metabolic reactions of existing drugs and also covert toxic compounds like pollutant to less or more toxic products. Enzymes can also be efficient in the remediation of pollutants.8–11 Thus, biocatalytic transformations have garnered lots of attention in organic and pharmaceutical syntheses, pollutant remediation, and metabolism studies.


image file: d3cs00461a-s1.tif
Scheme 1 Ribbon structure of human cytochrome P450 (CYPE14), identifying the iron heme cofactor that is central for all its biosynthetic reactions, and chart showing the relative importance of various liver enzymes in metabolizing known drugs. Reproduced from Anal. Chem., 2016, 88, 4584–4599 with permission.26 Copyright, 2016, ACS publication.

In this review, we summarize recent advances of bioelectrocatalytic transformations of organic compounds using Cyt P450s and peroxidases. Cyt P450s are an extensive family of enzymes in mammals, bacteria and plants.12,13 They have an iron heme cofactor (Scheme 1) imbedded within their polypeptide structures as the mediation center of all their catalyzed oxidation reactions (see Scheme 2). As mentioned, Cyt P450s metabolize drugs and other chemicals in humans, but in some cases produce toxic products.7,14–19 Cyt P450s are biocatalysts for C–H hydroxylation, N- and S-oxidation, O-, N- and S-dealkylation, C–C bond cleavage, C[double bond, length as m-dash]C double bond epoxidation, aromatic coupling, as well as more unusual reactions.20,21 They are effective for synthesis of drugs, drug metabolites,22–27 bioremediation,28–30 and development of biosensors.31–33


image file: d3cs00461a-s2.tif
Scheme 2 Pathway for mammalian Cyt P450-catalyzed reactions.8,13,35,41,42 The ferric iron-heme (P-Fe) Cyt P450 (1) first binds substrate RH to eliminate water (2). Next, 2 is reduced to 3 by NADPH-dependent reductase CPR to yield ferrous heme Cyt P450 3, which then binds dioxygen to form ferryl superoxy complex 4. This complex is reduced by NADPH reductase CPR to yield 5, which is then protonated to 6, which may also be generated by peroxide via the reversible peroxide shunt. Protonation of 6 and dioxygen cleavage lead to active heme-iron(IV)-oxo radical cation (7), considered the key reactive intermediate that transfers oxygen to the bound substrate to yield the product ROH. Exposure of ferrous form 3 to carbon monoxide (CO) yields P-Fe(II)-CO complex 9, which absorbs 450 nm light, which gave these enzymes the name Cyt P450s. Abbreviation: CPR, cytochrome 450 reductase.

Cyt P450s as monooxygenases incorporate one oxygen atom site-selectively into a reactive substrate.23,34 In humans Cyt P450 reductases (CPRs) accept electrons from nicotinamide adenine dinucleotide (NADH) or nicotinamide adenine dinucleotide phosphate (NADPH).35 In nature, CPR transfers two electrons from NADPH to Cyt P450s (Scheme 2). Other redox partners also exist for Cyt P450s,23,36e.g., bacterial and mitochondrial systems with NADPH-ferredoxin reductase and ferredoxin partners. Cyt P450s also have alternative activation pathways utilizing hydrogen peroxide (called the H2O2 shunt) to facilitate oxidative reactions.37,38 All these enzymatic reactions feature reduction of Fe(III) to Fe(II) that reacts with oxygen or H2O2 to yield reactive Fe(IV)[double bond, length as m-dash]O species to oxidize substrates.39 Cyt P450 oxidations have also been discussed in terms of heme iron–oxygen enzyme forms called compound I (P˙+Fe(IV)[double bond, length as m-dash]O), using terminology from peroxidases.40

Peroxidases use hydrogen peroxide (H2O2) as the oxygen source. The iron heme prosthetic group typically binds with a histidine residue that acts a proximal ligand. The catalytic cycle of all peroxidases is shown below and is similar to the H2O2 shunt process of Cyt P450s (eqn (1)–(3)). Peroxidases get activated by two-electron reduction of H2O2 to yield high valent Fe(IV)[double bond, length as m-dash]O species.

 
PFe(III) + H2O2 → PFe(III)–O–O–H → P˙+Fe(IV)[double bond, length as m-dash]O (Compound I) + H2O(1)
 
+Fe(IV)[double bond, length as m-dash]O + RH → Fe(IV)[double bond, length as m-dash]O (Compound II) + R˙ + H+(2)
 
PFe(IV)[double bond, length as m-dash]O + RH + H+ → PFe(III) + H2O + R˙(3)
Initially, peroxidase PFe(III) is oxidized by H2O2 to produce the two-electron oxidized intermediate (compound I, eqn (1)) compound I (P˙+Fe(IV)[double bond, length as m-dash]O).43–48 Compound I oxidizes a substrate to yield ferryl-oxy intermediate compound II (Fe(IV)[double bond, length as m-dash]O, eqn (2)), which is reduced back to the Fe(III) by oxidizing another substrate molecule (eqn (3)).44,49,50 In organisms, peroxidases destroy excess H2O2 and reactive oxygen species (ROS) as well as oxidize numerous compounds. Several types of peroxidases, e.g. lignin peroxidase,47,48,51–56 horseradish peroxidase (HRP),46,57–60 chloroperoxidase (CPO),61–64 manganese peroxidase,65,66 and haloperoxidases.67–69 have been used extensively in biocatalytic oxidations (e.g., the oxidation of alkanes to alcohols) or chromogenic reactions in chemiluminescence.70 While oxidation of these enzymes by H2O2 is required, their inactivation by H2O2 may limit their use in industry.71–73 Thus, recent studies have focused on in situ production of H2O2 that potentially overcomes the overoxidation issue.74

1.1 Classification of Cyt P450s and peroxidases

Cyt P450s are commonly divided into two classes. Bacterial and mitochondrial P450s are in Class I, and use a two-component electron transfer system containing a ferredoxin reductase and iron–sulfur protein (ferredoxin). Ferredoxins are acidic, low molecular weight, soluble proteins including iron-sulfur clusters, e.g. [2Fe-2S], [3Fe-4S] and [4Fe-4S], and a 7 Fe ferredoxin with both a [3Fe-4S] and a [4Fe-4S] cluster. The electron transfer pathway involves NAD(P)H donating an electron to ferredoxin reductase, which subsequently transfers to ferredoxin, and then to the Cyt P450.75–81 Mammalian microsomal Cyt P450s are called Class II enzymes and the electron source is NADPH Cyt P450 reductase.77–79 The generally accepted electron transfer and reductant oxidation pathway is presented in Scheme 2.77,82,83

While the short notation for these enzymes is Cyt P450s, specific enzyme notations are used for individual Cyt P450s, e.g., CYP3A4, CYP1A2, CYP2E9, CYP2B6, CYP2E1, CYP2C8, CYP2C19 and CYP2D6, etc.84 In those notations, CYP represents the superfamily followed by a number indicating the gene family, the capital letter after the first number represents the subfamily, and the last number denotes the individual gene.85 The use of italics in the name usually denotes a gene, e.g., CYP2E1 gene (i.e., the gene encoding the enzyme CYP2E1).86 Cyt P450s can be found in every class of organism. In humans, there are 18 families of mammalian Cyt P450s. Many of these enzymes are associated with specific drug metabolism and xenobiotic reactions to oxidize drugs/toxins.87,88 For example, CYP2E1 metabolizes paracetamol (acetaminophen) into N-acetyl-p-benzoquinoneimine.89 CYP2B6 metabolizes cyclophosphamide into 4-hydroxy cyclophosphamide.90 CYP2A6, CYP2B6, CYP3A4, CYP3A5, CYP2C9, CYP2C18, CYP2C19, show oxazaphosphorine 4-hydroxylation activity. High activity for 4-hydroxylation of cyclophosphamide is demonstrated by CYP2B6, and for 4-hydroxylation of ifosfamide by CYP3A4.91 The numerous androgen-metabolizing Cyt P450s (CYP3A4, CYP3A5, CYP3A43 and CYP2B6) and CYP enzymes (CYP27B1, CYP27A1, CYP24A1, CYP3A4, CYP2J2, CYP2R1) are necessary for vitamin D metabolism.85,92–94 Human CYP2B6, CYP2C19, CYP2D6, CYP2C9 and CYP3A4 metabolize 75–90% of existing phase I drugs. Among these, CYP3A4 is responsible for biotransformation of a very large number (close to 50%) of drugs.84,95,96

Peroxidase enzymes are found in plants, animal, and microorganism tissues. They are classified into three classes based on source.97–100 Class I-intracellular prokaryotic peroxidases include catalase, ascorbate peroxidase and cytochrome c peroxidase. Class II-extracellular fungal peroxidases contain manganese peroxidase, lignin peroxidase and versatile peroxidase, and class III-secretory plant peroxidases comprise horseradish root peroxidase (HRP), wheat peroxidase (WP), lignin-forming tobacco peroxidase (TOPA), tomato peroxidase (TMP) and barley peroxidase (BP).98,101 Peroxidases can use many types of peroxides as electron acceptors to catalyze oxidative reactions.

1.2 Synthetic scope of this review

Diverse applications of Cyt P450s and peroxidases, including biosensors, drug metabolism, synthesis reactions, catalysis, drug interactions and conversion have been reviewed by a number of research groups.7,31,32,102–108 We address below the use of Cyt P450s and peroxidases for electroorganic synthesis where no additional electron donors are required (see below). We organize the reactions in the text below from a synthetic viewpoint, based on the types of chemistry, e.g., C–H activation, S-oxidation, epoxidation, N-hydroxylation, oxidation of alcohols and oxidative N-, and O-dealkylation. There are four major goals of this review article,

(1) To provide a comprehensive overview of the status of Cyt P450s and peroxidases in bioelectrocatalysis, including suitability of electrode materials and immobilization methods.

(2) To identify reaction conditions required for high catalytic activity of these enzymes on electrode surfaces.

(3) To review recent bioelectrocatalytic applications of Cyt P450s and peroxidases for oxidative chemistry.

(4) To describe novel approaches used in bioelectrocatalysis with these enzymes.

2. Combining electrochemistry with biocatalysis

2.1 Rationales for electrochemical activation of heme enzymes

As shown in Scheme 2, two electrons are delivered sequentially to the iron heme in the complex Cyt P450 catalytic cycle. The first one reduces Fe(III) to Fe(II) in the heme group, and Fe(II) then reacts and binds with oxygen to yield Fe(III)–O–O˙. The second electron reduces this species to Fe(III)–O2, which eventually leads to the Fe(IV)[double bond, length as m-dash]O species that oxidize the reactant RH (Scheme 2).109 Cyt P450s are capable of oxidation or hydroxylation of non-activated carbon atoms, i.e., oxidation of C–H bonds. The use of NADPH as an electron source and cytochrome P450 reductases (CRP) as in the natural catalytic cycle (Scheme 2) can serve as an multifaceted biocatalytic system with significant industrial potential, but also carries practical issues that need to be solved.102,110–112 On one hand, the complex electron transfer mechanism of Cyt P450s in solution requires the use of expensive reductants such as NADH or NADPH. The monotonic eukaryotic Cyt P450 reductases, and bacterial redox partner system are more distinct but a bit complicated to apply to solution-based syntheses.20,113,114 On the other hand, although being difficult to handle, CRP that is necessary to the direct electron transfer with Cyt P450s presents in commercially available Cyt P450 preparations known as microsomes (or baculosomes) and supersomes. The latter constructs feature specific single Cyt P450's (e.g., CYP3A4) and include CRP in a predominantly lipid matrix. Microsomes are derived from specific organs and contain a collection of Cyt P450s, CPR, and other proteins in a lipid matrix. Cyt P450s catalyze regio- and stereo-selective activation of C–H bonds in organic reactants under optimal conditions.78,115 This makes them suitable as biocatalysts for the synthesis of pharmaceuticals and other fine chemicals where control of stereochemistry is absolutely important and necessary.

There have been extensive efforts to overcome the limitations of Cyt P450s in the past.116,117 Among those, electrochemical activation can improve performance of Cyt P450s without the use of additional chemicals. Replacement of biological electron transfers by electrochemical processes where electrons are transferred directly between electrodes and enzymes, or enzymatic systems offers a promising approach to more feasible, cost-effective, and simplified pathways. The driving force for enzyme-based reactions can be regulated by the potential applied on the working electrode in the electrochemical cell under control of the user.118 Immobilization of enzymes on working electrodes is an important tool to produce or generate the drug metabolites and other products with high yield and high selectivity for a fast toxicity screening.116,119–127 Electrochemical techniques can also scale up the generation of metabolites in sufficient amounts for investigation of metabolite or structural characterization. However, activation has often been driven by the H2O2 shunt, and not by the accepted biological pathway in Scheme 2.116 Nonetheless, electrochemical methods have many advantages for production of drug metabolites and other products, including: (i) simple approach; (ii) mild conditions in buffer solutions; (iii) avoid need for NADPH or NADH; and (iv) immobilization of enzymes on electrodes that greatly improves stability, and minimizes the amount of enzyme required.94,128,129

Cyclic voltammetry (CV) can be used to examine fundamental features of electrochemical transformations of enzymes. For example the standard heterogeneous electron transfer rate constant ks for Cyt P450 reductions in thin layer-by-layer (LbL) polyion films (see Scheme 4 and associated discussion) on pyrolytic graphite (PG) electrodes was shown to depend on the heme iron spin state.130 CVs of bacterial Cyt P450cam and human Cyt P450 2E1 in LbL film at pH 7.0 are shown in Fig. 1. These are background subtracted CVs of poly(ethyleneimine) (PEI)/(poly(sodium 4-styrene sulfonate) (PSS)/P450 2E1)4 and PSS(/PEI/P450cam)4 films on PG electrodes. The peak current was proportional to scan rate as predicted for surface bound species. In anaerobic buffers at pH ∼7, human Cyt P450cam with low spin iron gave a 40-fold higher electron transfer rate constant (ks, 95 s−1) for reduction than the high spin human Cyt P450 1A2 (2.3 s−1).130 The mixed spin CYP2E1 has an intermediate value of ks for reduction. In addition, comparison of experimental CVs with digital simulations was consistent with faster oxidation rates of the Fe(II) forms than reduction rates of Fe(III) forms. This suggested that Cyt P450s retain their enzymatic activity on the electrode where an underlying chemical conversion scheme involving oxidized and reduced forms of Cyt P450s underwent rapid conformational equilibria coupled with their electron transfer processes.130 The rate constant for chemical oxidation of Fe(III)-Cyt P450s in these films by t-BuOOH to active Fe(IV)[double bond, length as m-dash]O Cyt P450s obtained by rotating disk voltammetry (RDV) also showed dependence on spin state. Therefore, those enzymes can maintain native structures and enzyme activities in the LbL films, that is fundamentally important for electrochemical activation of heme enzymes.


image file: d3cs00461a-f1.tif
Fig. 1 Background subtracted LbL films of enzymes with polyions on a PG electrode in anaerobic 50 mM phosphate buffer with 0.1 M sodium chloride, pH 7: (A) Cyt P450 2E1/polyion, and (B) Cyt P450cam/polyion. Reproduced from J. Am. Chem. Soc., 2009, 131, 16215–16224 with permission.130 Copyright, 2009, ACS Publications.

2.2 Activation of enzymes in electrochemistry

As suggested above, immobilization of Cyt P450s or peroxidases on an electrode surface does not stop their binding of substrates and their further catalysis. Immobilization on electrodes can also stabilize enzymes, minimize the amount of enzymes (e.g., down to nM) needed, and eliminate the use of electron donors. The enzyme co-factors or initial acceptors like CPR can accept electrons directly from the electrode that can act as an electron source for P450s.28–31 In bioelectrocatalysis, enzymes can mimic their natural pathway for Cyt P450-catalyzed oxidations on electrodes as demonstrated in the thin LbL films made from polyions, and Cyt P450 CPR microsomes containing pure Cyt P450.131 These stable films consisted of thin layers of the polyions poly(diallyldimethylammonium chloride) (PDDA) and PSS as six alternating layers with CYP2A1 microsomes (containing CPR), denoted as PDDA/PSS(/CYP1A2/CPR)6 films on PG electrodes. Comparing voltammetry with CYP1A2 and CPR films alone showed that electron flow proceeded from the electrode to CPR and then to CYP2A1 in the CYP1A2/CPR films, very similar to the biological pathway in Scheme 2. The background subtracted CVs of the CPR/CYP2A1 films are characteristic of the electrochemical reduction-oxidation reactions (Fig. 2A). The films of CPR/CYP2A1 showed a peak at −230 mV vs. normal hydrogen electrode (NHE) and there was no peak potential shift when carbon monoxide was added. The electron transfer rate constant (ks) was ∼40 s−1, similar to films with only CPR or CPR/cytochrome b5 present. Fig. 2B shows that CYP1A2 alone has a peak potential −90 mV vs. NHE. There was a ∼35 mV shift with addition of CO characteristic for Cyt P450s, and 20-fold smaller ks values than films with CRP and CYP2A1. Similar results were observed when CYP2E1 was used in the same experiments. Those results support the sequential transfer of electrons from electrode to CPR to Cyt P450s when both are present in the film. Electrons are not directly transferred to Cyt P450, when CPR and Cyt P450s are both present.
image file: d3cs00461a-f2.tif
Fig. 2 Experimental and simulated voltammetry of CYP1A2 + CPR microsome films on pyrolytic graphite electrodes. (A) Background subtracted voltammograms of (a) CPR films at 0.3 V s−1 without CYP and (b)–(d) with CYP1A2 film with CPR films at (b) 0.1, (c) 0.2, and (d) 0.3 V s−1 scan rate. (B) Background-subtracted voltammograms of (a) polyion film and (b) and (c) CYP1A2 film without CPR at (b) 0.1 and (c) 0.2 V s−1 scan rate. (C) Digitally simulated theoretical voltammograms related to (a) reversible electron transfer at CPR film at 0.3 V s−1 scan rate and (b)–(d) ErCEo-model and parameters in Scheme 1 for CYP1A2 and CPR films at (b) 0.1, (c) 0.2, (c) 0.3 V s−1 scan rate, demonstrating the excellent agreement with experimental voltammograms in (A). (D) Effect of scan rate on reduction (red diamonds) and oxidation (blue circles) peak potential for CYP1A2/CPR films plotted with theoretical peak potentials (lines) simulated by using the ErCEo model. Reproduced from J. Am. Chem. Soc., 2011, 133, 1459–1465 with permission.131 Copyright, 2011, American Chemical Society.

Fig. 2C shows CV digital simulations of the model featuring electron transfer first to CPR then to Cyt P450 when both are present in the film. Simulations showed good agreement with experimental CV peak potentials and shapes. The simulation model features electron transfer from CPR to Cyt P450 and electrochemical–chemical–electrochemical (ErCEo) pathway (eqn (4)–(6)), where Eo is electrochemical oxidation and Er is reduction of CPR. Initially, CPR accepts an electron characterized by electrochemical rate constant ks,red (eqn (4)). Eqn (5) represents redox equilibrium between the CPRred and Cyt P450-FeIII to produce CPRox and Cyt P450-FeII, followed by oxidation of CPRred (eqn (6)).

 
CPRox + e → CPRredE° = −230 mV vs. NHE; ks,red = 0.4 s−1(4)
 
CPRred + P450-FeIII ⇌ CPRox + P450-FeIIkb ≥ 5kf s−1(5)
 
CPRred → CPRox + e E° = −230 mV vs. NHE; ks,ox = 40 s−1(6)

Electrons flowing from electrode to CPR to Cyt P450 was confirmed by observing the catalytic limiting current from RDV for oxidation of 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone (NNK) to 4-hydroxy-1-(3-pyridyl)-1-butanone (Fig. 3) and by confirming the product formed in this reaction.131 The Cyt P450 on the electrode surface increased the plateau current with increasing the concentration of NNK (Fig. 3). The increment in the current demonstrated that the oxidation of NNK was catalyzed by Cyt P450/CPR film.


image file: d3cs00461a-f3.tif
Fig. 3 Rotating disk voltammograms of Cyt P450s/CPR film with increasing the concentration of NNK in pH 7.0 buffer + 0.1 M NaCl at 1000 rpm. Regenerated from J. Am. Chem. Soc., 2011, 133, 1459–1465 with permission.131 Copyright, 2011, ACS publications.

In a related work, the Cyt P450 3A4 supersomes with CPR were immobilized on colloidal Au coated graphene electrode through electrostatic interactions. Electron transfer from electrode to CPR to CYP3A4 was also observed in this system.127 The electrochemical catalytic behavior of CYP 3A4/CPR was examined by RDV (1000 rpm) in aerobic pH 7.4 phosphate buffer (0.1 M). The reduction peak current increased with increasing the concentration of nifedipine from 0 to 7.44 μM at −0.5 V (Fig. 4).


image file: d3cs00461a-f4.tif
Fig. 4 The RDV (1000 rpm) of CYP3A4/CPR films with increasing the concentration of nifedipine in aerobic 0.1 M phosphate buffer (pH = 7.4) (a) 0, (b) 1.24, (c) 2.48, (d) 3.72, (e) 4.96, (f) 6.2 and (g) 7.44 μM. Regenerated from RSC Adv., 2012, 2, 12844 with the permission.127 Copyright, 2012, RSC.

As described above, the bioelectrocatalytic pathway in Cyt P450s including CPR on electrode surfaces is the same as the natural human mechanism, featuring the initial electron transfer to microsomal reductase CPR then from CPR to the Cyt P450-heme-iron. We found this to be true as well for both microsomes and supersomes, already containing CPRs, in LbL films on electrodes.132,133 While prior work revealed that electrodes can serve as an alternative source of electrons in place of NADPH or NADH for catalysis by Cyt P450s, these systems required peroxide to activate the enzyme for electrocatalytic syntheses oxidation by the H2O2 shunt process.95

Bioelectrocatalytic activation Cyt P450s supersomes on electrode surfaces in LbL films of DNA has also been developed as a general approach to monitor rates of DNA damage by metabolites.23–25 The general procedure employs microfluidic arrays featuring individual electrode wells coated with LbL films made with different CYP microsomes, ruthenium polymer Ru(bpy)2-PVP (PVP = polyvinylpyridine) and ds-DNA (ds = double stranded). Reactant flows into the array under an applied oxidation potential that activates the CYP microsomes for oxidation of the reactant to form metabolites that can react with DNA to cause partial ds-DNA unwinding. The relative rates of DNA damage by different Cyt P450s are subsequently detected by electrochemiluminescence (ECL) activated at about +1 V vs. SCE and detected with a CDC camera in a dark box.23–25 Guanines in the partly unraveled DNA act as ECL coreactants that are sensitive to the amount of DNA damage since they are now more accessible to react with the Ru(bpy)2PVP ECL dye. Specific samples of this approach to monitor chemical toxicity are described later in this review.

Additionally, non-natural electrochemical mediators can be used to shuttle electrons to enzymes. For example, Co(II) sepulchrate trichloride can replace NADPH to activate recombinant fusion enzyme rFP450 for the ω-hydroxylation of lauric acid in solution.134 1,1′-Dicarboxycobaltocene can replace NADPH to activate P450BM3 that catalyzes hydroxylation of lauric acid.135 Coupling with non-natural electrochemical mediators can extend the practical applicability of electrochemical mediation of electron transfer to P450s as a catalyst for regio-selectivity and stereo-selectivity.

3. Challenges in enzyme electrocatalysis and their solutions

Cofactors are nonproteinogenic molecules that are needed for catalytic activity of enzymes which usually bind to enzymes through covalent or electrostatic interactions and are known as prosthetic groups. A broad variety of cofactors are well-known, and may be organic molecules, inorganic ions or metal complexes. In the case of covalent binding, the enzyme is permanently bound to the cofactors. The coenzymes are modified during catalytic reactions by electron transfer or other reactions with the reactant or substrate. Cofactors that are organic molecules are called coenzymes. Regeneration of coenzymes is a key challenge for use in some catalytic reactions.136 The heme cofactor features a porphyrin ring with a central iron ion. It is found in all known forms of life (bacteria to humans).137,138 Iron heme is the prosthetic group in hemoproteins such as myoglobin, hemoglobin, peroxidase, cytochromes, catalase, albumin, prion protein, guanylate cyclase and nitric oxide synthase (Scheme 3).139–141 The central iron atom binds the oxygen and transfers it in various forms to small reactant molecules. There are various unbound cofactors used for enzymatic synthesis such as NADH, adenosine triphosphate (ATP), flavin mononucleotide (FMN), and thiamine diphosphate (ThDP).136 In the Cyt P450 catalytic cycle (Scheme 2), NADH or NADPH transfer electrons in the enzymatic pathways and must be regenerated to continue the catalytic cycle. Regeneration of the co-enzymes can be energetically expensive and may require additional co-substrate or additional enzymes.142–146 Co-enzymes may also present challenges in terms of availability, and recycling.147
image file: d3cs00461a-s3.tif
Scheme 3 Iron heme that serves as prosthetic group in Cyt P450s and peroxidases.

Mammalian Cyt P450 enzymes typically accept two electrons originally from reducing agent NADPH that transfers them to flavoprotein NADPH-Cyt P450 oxidoreductase (CPR), the immediate electron donor for Cyt P450s (Scheme 2).148 However, all reducing agents are not equally effective at donating electrons to Cyt P450s. Hydrogen peroxide or other peroxides may also serve as sources of oxygen for Cyt P450 to catalyze the substrate hydroxylation. CPR can also produce reactive oxygen species (ROS) resulting in NADPH consumption by microsomal P450. These reactive species can cause oxidative damage to enzymes. Thus, it is important to control and minimize generation of reactive oxygen species.149 This oxidative damage can be controlled by using the more specific electron donor and antioxidants to work alongside Cyt P450 enzymes.150

Instability of enzymes can degrade catalytic performance and shorten lifetimes. Numerous factors affect the stability of enzymes, including pH, temperature, and the presence of substrates or inhibitors. Thermostable P450 enzymes can play a significant role in synthesis of the valuable organic compounds. Thermophilic enzymes, such as CYP116B29, CYP116B46, CYP116B64, CYP116B65, CYP119, CYP119A1, CYP119A2, CYP154H1, CYP174A1, CYP175A1, CYP174A1, and CYP231A2 can be used at higher temperature than normal for enzymes.151–153 The thermal stability of these enzymes makes them attractive for industrial use. For example, a CYP119/sol-gel film remained stable up to 60 °C, but the reduction current decreased to ∼45% as compared to that at 30 °C. An enzyme electrode constructed by using LbL films with PDDA and PSS improved thermal stability. CYP119/PDDA/PSS films retained electrochemical activity up to 93% at 80 °C (Fig. 5).154 Higher temperatures can increase the reaction rates significantly, which is of key important for industrial applications. When CYP119 was immobilized in films of dimethyldidodecylammonium bromide and PSS (DDAB/PSS) and used for CCl4 dehalogenation, the formation rate of methane increased by 35-fold from 25 °C to 55 °C. Other polymers like poly(ethylene oxide) (PEO), when chemically grafted on enzymes, can provide remarkable thermal stability. PEO-modified CYP119A2 was active up to 120 °C in 0.5 M of KCl with a small potential shift of redox potential of ∼14 mV between 25 °C and 120 °C.155 These results show that stability and electrochemical activity of enzymes at temperatures much higher than for the free native enzyme in solution can be improved by making conjugates with suitable polymers as elaborated in more detail below.


image file: d3cs00461a-f5.tif
Fig. 5 Catalytic conversion of CCl4 into CH4 using CYP119/DDAB/PSS films (left). The effect of temperature on catalytic dehalogenation of CCl4. Linear sweep voltammograms of CYP119/DDAB/PSS in the presence of CCl4 at 25 °C, 55 °C, and 75 °C (right). Reproduced from, J. Am. Chem. Soc., 2004, 126, 8632–8633, with permission.154 Copyright, 2004, ACS publications.

3.1 Enzyme film formation and electrode materials

The immobilization of Cyt P450s and peroxidases on electrodes can remarkably improve stability and activity. Immobilization of enzymes can be done by many different methods,106,112,156–162e.g., covalently bonding, SAMs, lipid bilayer films, and LbL films. Immobilization often includes the following key steps,

(i) Clean or activate the electrode by mechanical polishing or electrochemical potential cycling. For some films, roughening the electrodes and introducing surface charges chemically can be advantageous for film stability.163

(ii) Functionalize electrode surfaces with enzymes via drop casting, self-assembled monolayers (SAMs), incorporation into detergent or lipid films, LbL film growth, or chemical attachment.106,112,164–166

(iii) Age electrodes at 4 °C in a constant humidity chamber to avoid complete dehydration and remove weakly adsorbed enzymes from the surface by washing.167–169

3.1.1 Covalent immobilization of enzymes. Covalent immobilization of enzymes can be used to immobilize the enzyme on the electrode surface. Functionalizations used in attachment of the enzyme include amidization by –NH2 and –COOH coupling catalyzed by N-((3-dimethylamino)propyl)-N-ethyl carbodiimide hydrochloride (EDC)-N-hydroxysulfosuccinimide (NHSS), and thiol–ene click chemistry using cystine (see Table 1). Bioconjugation chemistry using EDC-NHSS cross-linking can immobilize enzymes on electrode surfaces. NHSS is an effective way to stabilize the reaction intermediate and improve coupling reactions with primary amines. NH2 and COOH groups can be added onto electrodes through surface ligand chemistry (e.g., metal–thiolate coordination on an Au electrode) or physically adsorbed with polyions (e.g., oxidized carbon bound with positively charged polyamines). Surface NH2 or COOH can now react with enzymes through EDC-NHSS amidization.170–172 Afterwards, chemically bound enzymes can be partially dried in a constant humidity chamber.170,173–176 Using human CYPs 3A4, 2D6 and 2C9 in an electrochemical array, Gilardi et al. investigated how immobilization would impact the binding affinity (with Michaelis constant KM as an indicator) of substrates in a high throughput microtiter-based platform.173 Cyt P450s were immobilized on hydroxyoctanethiol and 10-carboxydecanethiol (1[thin space (1/6-em)]:[thin space (1/6-em)]1) self-assembled monolayer (SAM) on a gold (Au) electrode. EDC-NHSS coupled carboxylic groups in the SAM film with amides on Cyt P450s. Michaelis–Menten analysis was done using the current measured at −390 mV with 30 known drugs. KM values were slightly better or on par with those measured for Cyt P450 microsomes in solution. Also, nanomaterials coupled with enzymes can improve their electrochemical activation. For example, CYP2B6 on a glassy carbon electrode modified with ZrO2 nanoparticles and platinum-PLL (Pt-PLL) showed reversible electron transfer between heme iron of CYP2B6 and electrode at −0.449 V at pH 7.4 in anaerobic solution.177KM was lowered by two orders of magnitude as compared to CYP3A4 in microsomes.
Table 1 Immobilization techniques to prepare enzymatic electrodes
Method Enzyme Crosslinker and mixture Incubation time Electrode Enzyme activity Stability Enzyme leakage Ref.
Covalent binding Formate dehydrogenase EDC and NHS 30 min for enzymes, and 90 min for EDC/NHS (i) 4-ATP SAMs; (ii) 4-nitrophenyl electrochemically grafted on Au electrode Low High No 188
P450 BM3 Dendritic mesoporous silica with –OH and –NH2 12 h CS/GCE 189
CYP101 N-(4-Caboxyphenyl) maleimide and SWCNT 30 min at 4 °C GCE 190
P450 BM3 (i) Cystamine-N-succinimidyl 3-maleimidopropionate; (ii) dithio-bismaleimidoethane 1 h at 4 °C Au(111) surfaces 191
CYP 2B4 EDC/NHS 90 min at R.T. Carbon and Au SPE 170
P450 2E1, cysteine mutant MUT261 or MUT268 Dithio-bismaleimidoethane (DTME) 2 h for DTME Au 124
Physical adsorption P450 BM3 30 min, at 4 °C ITO Moderate Low Yes 192
P450 3A4 12 h, at 4 °C DDAB/SPE or MWCNT/SPE 193
P450 3A4, 2B4, 1A2, 11A1 (P450scc), and 51b1 12 h, at 4 °C, Humid chamber Au nanoparticles-DDAB/PG 157
CYP2B6 Au-chitosan mixed with CYP2B6 12 h at 4 °C GCE 158
CYP2C9 ITO and CS mixed with enzyme GCE 194
Mb, HRP or hemin Mb, HRP or hemin mixed with DDAB Covered with vial and dry overnight (∼12 h) PG 181
LbL films CYP1A2 PLL or poly-L-ornithine 15 min for polyelectrolyte, 30 min for CYP1A2 Quartz crystal Moderate High No 184
CYP1A2 PDDA/PSS PG 131
P4502E1, P450cam, P450 1A2, catalase and Mb PEI/PSS PG 130
Mb and P450cam PEI, PDDA and PSS 15 min for polyions Au 187
NADPH-cytochrome P450reductase PDDA and PSS 20 min for polyions and 1 h for enzymes PG 195
CYP3A4 PDDA GCE 125


3.1.2 Incorporation of enzymes through non-covalent interactions. Physisorption has been used frequently to prepare enzyme-coated electrodes due to its simplicity. The enzyme is physically adsorbed on the electrode surface. Adsorption occurs via non-covalent forces including electrostatic interaction, van der Waals, hydrophobic interaction, and hydrogen bonding.178 Although the activity of the enzyme is largely retained as compared to covalently grafted enzymes, the loosely bound enzymes can desorb under reaction conditions, and must be avoided. A comparison of immobilization methods is summarized in Table 1.

Films of insoluble surfactants can accommodate a wide variety of enzymes and heme proteins and have enabled reversible electrochemical reduction of FeIIIheme cofactors. These detergent-protein films are made by mixing dispersions of water, insoluble surfactant, and protein, drop casting them on the electrode, and drying before use. They feature multiple layers of bilayers formed with insoluble double chain surfactants such as didodecyldimethylammonium bromide (DDAB) and phospholipids, with the proteins residing between the detergent bilayers. These interlayer regions also contain considerable amounts of water that is available to hydrate the enzymes.179–181 Direct immobilization of Cyt P450s has been achieved in electrode-supported lipid bilayer membranes, similar to those developed for hydrophobic membrane enzymes.182,183 For example, CYP2C9, and its polymorphic modification CYP2C9*2 and CYP2C9*3 could be immobilized on DDAB modified screen printed electrodes (SPE).156 Well-defined reduction and oxidation peaks of the FeIIIheme were obtained in CVs. The midpoint potential (Emid) for CYP2C9, CYP2C9*2 and CYP2C9*3 were observed −0.318, −0.324 and −0.318 V vs. Ag/AgCl, respectively. In aerobic conditions, the reduction potential of CYP2C9*2 and CYP2C9*3 shifted towards the positive potential in comparison to CYP2C9 with ∼50 mV vs. Ag/AgCl. The shift in the potential demonstrated easier electron transfer to the heme iron in CYP2C9*2 and CYP2C9*3. Immobilized CYP2C9, CYP2C9*2 and CYP2C9*3 were all active for bioelectrocatalytic 4-hydroxylation of diclofenac in the presence of antioxidants mexidol or taurine at −0.55 V. In comparison with the microsomal systems, the Michaelis constants (KM) of those three forms of CYP2C9 increased by 10–20 times, suggesting that best structural conformation of CYP2C9 may vary to slightly limit the substrate binding on the electrode.

LbL films first developed by Lvov and Decher have been used to grow stable enzyme films on electrode surfaces.131,184–186 In this method, also mentioned previously, multiple layers of polyions or nanoparticles are assembled by adsorbing one layer at a time with alternate layers of oppositely charged enzymes, microsomes or supersomes (Scheme 4). These thin films are strongly absorbed onto the electrode surface, usually by placing single drops of each adsorbate solution on the electrode for 15–20 min. Weakly adsorbed species are washed off the electrode after each adsorption–equilibration step and dried in a stream of nitrogen before application of the next layer.


image file: d3cs00461a-s4.tif
Scheme 4 Illustration of layer-by-layer (LbL) film formation of enzymes and polyions.

LbL methods provide high stability to the enzyme film without desorption. Forces between layers are primarily electrostatic in nature, but may include hydrophobic interactions, coordination, hydrogen bonding, or even covalent bonding.165,185,186 The enzyme attached to the electrode surface through the consecutive addition of negatively or positively charged polyions or layer of the enzymes for example, polyions of PDDA, PSS, PEI, etc. (see Table 1), all of which provide multiple electrostatic attraction sites of bound enzymes to strongly hold them on electrodes. Lvov working in our group reported the first LbL film direct voltammetry of enzymes and redox proteins.187 Myoglobin (Mb) and bacterial Cyt P450cam were used in LbL films with DNA and polyions on an Au electrode coated with chemisorbed mecaptopropanesulfonic acid to provide stable, electroactive multi-layered films. Mb films were made by successive adsorption of PSS, DNA or PDDA. The Cyt P450cam film was grown with layers of PSS or PDDA. LbL films of Cyt P450s using microsomes or supersomes can be assembled on electrodes similarly, as discussed above (see Table 1.)131

3.2 Cells and electrode materials

Electrochemical cells and electrodes sizes used in synthesis need to be significantly different from those used for techniques such as voltammetry. In voltammetry, working electrodes are typically disks of 10 μm to several mm in diameter, and counter electrodes are typically a Pt wire. However, in constant potential electrolysis, the goal is to make a measurable amount of product, and
 
The rate of electrolysis = (const) × ekelectt(7)
where kelect is the synthetic rate constant and t is the reaction time. The reaction rate decreases exponentially with t as the decrease of the reactant concentration. The synthetic rate constant image file: d3cs00461a-t1.tif, where m0 is mass transfer coefficient, A is the electrode area, and V is the solution volume, respectively.196

The mass transfer coefficient m0 is calculated as image file: d3cs00461a-t2.tif, where δ0 is the Nernst diffusion layer thickness and D0 is the diffusion coefficient. More efficient stirring gives a smaller δ0, leading to faster mass transfer and synthetic rate. The concentration of the reactant at any given reaction time is C(t),

 
image file: d3cs00461a-t3.tif(8)
where CO is the initial reactant concentration, and the current It at an any given reaction time is given by
 
image file: d3cs00461a-t4.tif(9)
Thus, to obtain good synthetic yields in a high kinetics, the synthetic rate constant should be maximized by using a large electrode area to solution volume ratio (A/V), and ensuring efficient mass transport. If a large working electrode is used, a similar sized counter electrode should be used to keep current density on both electrodes similar. Symmetric placement of working and counter electrodes with the reference electrode tip close to the working electrode also helps to keep the potential constant across the entire working electrode.

The current I of the bioelectrocatalytic reaction can be measured during synthetic reaction to obtain current efficiency, which is the fraction of charge Q being used to produce the desired product. where

 
image file: d3cs00461a-t5.tif(10)
Current efficiency = Qr/Qt. Here, Qr is the charge needed to produce the amount of product formed obtained from Faraday's law.196

Binding of the enzymes on the working electrode depends to a certain extent on the electrode material and surface pretreatments. Several electrode materials including glassy carbon (GC), PG, graphitic carbon, carbon felt, carbon cloth, indium tin oxide (ITO), platinum, gold, reticulated carbon, and screen-printed carbon (SPE) have been used to immobilize enzymes.123,157,159,170,197–201 Au is often used for immobilization in smaller scale projects because it can covalently bind with sulfide groups (e.g., thiol, disulfides, thiolate, and thioesters), with the resulting Au–S–R leading to versatility in the structure of R. For example, R could be a charged group to electrostatically bind to an oppositely charged enzyme, or a reactive group that can be chemically coupled to an enzyme. However, Au and Pt are probably too expensive for industrial scale electrosynthesis. Among the many types of carbons, rough PG is particularly useful for film formation since it is high energy surface capable of adsorbing many different types of polyions. Polished glassy carbon has low surface energy and is not particularly good for film formation. High surface area materials such as carbon felt, carbon cloth and reticulated carbon are also advantageous due to the requirement for high A/V ratios (eqn (8)).

The biocatalytic activity of the electrode can sometimes be improved by using nanomaterials such as metal nanoparticles, conducting polymers, carbon nanomaterials, and thiolate compounds for protein-electrode attachment. When oxidized, these materials can provide functional groups to effectively interact with enzymes and shuttle electrons. The efficiency of the electrochemical synthesis can be improved by using nanostructured surface modifiers on the electrode.157,159,198–201 For example, multi-walled carbon nanotubes (MWCNTs) and DDAB on screen printed carbon electrodes (SPE) were compared for electrocatalysis with CYP3A4.193 For catalytic oxygen reduction, the cathodic current increased up to 3-times with CYP3A4/SWCNT/SPE in comparison to CYP3A4/DDAB/SPE. Additionally, CYP3A4 and gold nanoparticles (AuNPs) stabilized by DDAB films on SPE were used to investigate the effect of the B Vitamins (B1, B2, and B6) on the hydroxylation of diclofenac.202 Similarly, Cyt P450/carbon nanofiber (CNF) films were used on electrodes without a mediator layer and direct electron transfer between CNF and CYP3A4 was observed at −0.30 V vs. Ag/AgCl.199 High catalytic activity for oxygen reduction was found at CYP3A4/CNF compared to other carbon electrode nanomaterials like carbon black. Thus, catalytic activity of the enzyme electrode can often be improved by using optimal nanomaterials.

3.3 Activation of Cyt P450s

The heme reactive center of Cyt P450s is activated for biocatalytic reactions through the electron transfer in the presence of oxygen (see Scheme 2). In electrochemical reactions, electrons are transferred from electrodes instead of the natural electron donors NADPH or NADP.203 Direct electron transfer from electrode to the heme iron was reported for electrocatalysis using Cyt P450 as discussed previously.204 The reversible CV peaks for FeIII/II should be examined prior to any electrochemical catalysis. For example, CYP2C18 immobilized in an insoluble DDAB surfactant film on edge plane pyrolytic graphite (EPG) gave reversible CV peaks for FeIII/II. Fig. 6 demonstrates the reversible Fe(III) → Fe(II) redox couple (anodic peak current (ipa)/cathodic peak current (ipc) ∼ 1) of CYP2C18. In the presence of trace amounts of dioxygen, well-defined cathodic peaks with a slightly positive shift could be seen and were assigned to catalytic reduction to yield ferrous P450.205
image file: d3cs00461a-f6.tif
Fig. 6 Background subtracted CVs of CYP2C18 in presence (solid curve) and absence (broken curve) of 150 mM phenytoin at 100 mV s−1 scan rate (pH 8.0) showing reduction in upward direction. Reproduced from Electrochem. Commun., 2005, 7, 437–442 with permission.205 Copyright, 2005, Elsevier.

Electrocatalysis using Cyt P450s can be assessed using oxygen reduction. CYP27B1 immobilized on an EPG electrode in a DDAB film gave a significant increase in the cathodic current in the presence of oxygen due to electrocatalytic reduction of dioxygen to peroxide.206 The catalytic oxygen reduction current showed a significant increase along with the positive shift of potential when CYP101 was embedded in DDAB films, as compared to pure DDAB films. As mentioned in Section 3, LBL films of microsomes207 or supersomes can be electrochemically activated as the natural catalytic cycle involving electron donation first from electrode to CPR, then from CPR to Cyt P450s. Bioelectrochemical oxidation with heme enzymes features specificity and regioselectivity, but, electrochemical oxidation, in the absence of enzymes does not provide this advantage. Direct comparisons of electrochemical and bioelectrochemical reactions are summarized in Table 2. For example, electrochemical oxidation of diclofenac catalyzed by CNT and/or metal nanoparticles (Pt and Ru) shows more than one product (Table 2), while bioelectrocatalysis with CYP3A4 and CYP2C9 is highly regioselective and gives 4-hydroxydiclofenac as the only product. Likewise, the enzymatic electrooxidation of styrene, benzo[o]pyrene, and tramadol shows similar trends. With enzymes, the regioselective product is obtained; while, in the absence of enzymes, often more than one product are generated in electrooxidation. Therefore, bioelectrocatalysis provides high selectivity, reduces the electrolysis time, lowers electric energy input, and often produces targeted molecules with no isolation from side-products (see Table 2).

Table 2 Electrochemical and bioelectrochemical reaction comparison
Electrochemical method Bioelectrochemical method
Electrode Potential, electrolysis time Substrate Product (s) Product identification methods Ref. Electrode Electrolysis time Substrate Product (s) Product identification methods Ref.
CNT, Pt/CNT and Ru/CNT 1.5 V vs. SCE, 8 h Diclofenac Formic acid, acetic acid, oxalic acid, oxamic acid, maleic acid, malonic acid, piruvic acid, glioxilic acid, fumaric acid, glicolic acid, succinic acid, butiric acid, tartronic acid HPLC-MS, ESI-MS, ESI-MS/MS 223 CYP 3A4 + CPR, CYP 2C9 + CPR −0.6 V vs. Ag/AgCl, 45 min Diclofenac 4-Hydroxydiclofenac HPLC, MS 224
GF-CoS2/CoS 4.6–5.1 V, 4 h Styrene Benzaldehyde, styrene oxide, dibromide, β-bromostyrene, GC-MS, 1H-NMR 225 PSS-(CYP1A2-PSS)2 −0.6 V vs. SCE, 1 h Styrene Styrene epoxide, benzaldehyde GC 226
Pt electrode, 0.1 M tetraethylammonium perchlorate +1.3 V Benzo[o]pyrene 6-Acetoxybenzo[o]pyrene, and mixture of benzo[o]pyrene-quinones Fluorescence, and absorption spectroscopy 227 CYP1A1/PBA-NGO/GCE −0.48 V vs. SCE, 5 h Benzo[a]pyrene Benzo[a]pyrene-7,8-diol HPLC, GC-MS 228
Boron-doped diamond and Pt 10–30 mA, vs. Ag/AgCl, 1 h Tramadol N-Desmethyl- and O-desmethyltramadol as well as tramadol-N-oxide LC-MS, NMR 229 CYP2D6/PEI-RGO/GC −0.52 V vs. SCE, 2 h Tramadol O-Dimethyl-tramadol HPLC-MS, MS 230


Orientation of enzymes on the electrode surface may play a significant role in maintaining its activity. Enzymes covalently immobilized on electrodes in a single layer may be limited in terms of mobility and configurational dynamics by their orientation. The electron transfer efficiency between electroactive site and the electrode is controlled by the distance between the electrode and enzyme electroactive site. However, it is important to realize that orientation issues will be important mainly in cases where the enzyme has little or no mobility after immobilization, such as in covalent attachment to a solid electrode. In some films, enzymes retain significant mobility to make orientation a non-issue, such as in liquid crystal surfactant films, LbL polyion films, and hydrogel polymer films. In all of these films, enzymes are present in more than a monolayer and electrons transfer occurs predominantly by an electron hopping mechanism involving a reduced enzyme-oxidized enzyme reaction progressing through the film.208

Although mediated electron transfer using a molecular redox mediator is possible, direct electron transfer at the enzyme-electrode interface is more efficient with fewer steps and more efficient enzymatic activation process. Therefore, the configuration of heme enzymes in a monolayer, for example, needs to be orientated to minimize the distance between the heme center and electrode (within 14 Å) as a so-called “electroactive” configuration.209–211 HRP, for example, has a molecular radius of 30 Å212 that would block the direct electron transfer, since the heme sites buried in the protein frameworks may be too far away from the electrode. Previous literature suggested that the selectivity and stability of enzymes can also be controlled by varying the orientation of enzymes on the electrode surface.213 Although the orientation of enzymes can be significantly impacted by film formation condition, e.g., pH and ionic strength, there are several methods to control the orientation of enzymes on electrodes, including anchor chain molecular system,214–216 site-specific attachment,217 and antibody-based immobilization.218–222

Covalent attachment onto a specific site of enzymes may provide some degrees of control over the orientation of enzymes on the surface of electrode.213,231 The common strategy is to use the cysteine-based selective surface attachment, including cysteine-electrode binding232 (or metal–thiolate coordination) and thiol–ene click reaction.124,233 For example, CYP2E1 wild-type and cysteine mutants (MUT261 and MUT268) can be covalently immobilized on an Au electrode using the cross-linking chemistry of dithiobismaleimidoethane (DTME).124 DTME is a maleimide crosslinker for cysteines. This covalent linkage enables the binding of MUT261 and MUT268 in a way that orientation of the proximal electron transfer side of the heme is close to the electrode; while, the distal substrate-binding side is oriented towards the bulk solution for easy access of the substrate. Efficient electron transfer between heme and electrode can be achieved where MUT261 and MUT268 produced ∼2.5 and 2.0 times more products compared to that of wild-type CYP2E1. The site-specific attachment at residues 261 and 268 on proximal side of cysteine mutants are more efficient for electrocatalysis.

Site-specific modification on enzymes that can provide the oriented enzyme-electrode interaction is of increasing interest. Adding a short electrode-binding peptide at a specific terminal of enzymes can guide the binding direction of enzymes to the electrode and avoid the non-specifical orientation of enzymes.209 This method can provide the oriented immobilization of Cyt P450s.234 P450 BM3 C-terminally high specific peptide (HSP), as an example can be immobilized on an ITO electrode binding peptide (HSP) made of sequence (RTRHK)4. HSP has a profound effect on the protein orientation thus the electron transfer rate. P450 BM3 without HSP were highly active for de-ethylation of 7-ethoxycoumarin (7-EC) in solution (Fig. 7). When immobilized on the ITO, the orientation of enzymes induced by HSP improved the activity of P450 BM3 in either NADPH-mediated or direct electron transfer pathways. In case of direct electron transfer, P450 BM3 with HSP on ITO under a fixed orientation enhanced the activity of the enzymes up to 10-fold in comparison to that without no HSP (Fig. 7b–d), because random orientation of enzymes limited active center available for electron transfer.


image file: d3cs00461a-f7.tif
Fig. 7 7-Ethoxycoumarin (7-EC) activity assay on P450 BM3 specifically (HSP) and P450 BM3 unspecific (noHSP) immobilization, (a) dealkylation of 7- EC to 7-hydroxycoumarin (7-HC) by P450 BM3 mutant (A74G, F87V, L188Q). (b–d): Activity assay of noHSP and HSP on the production of 7-HC/pmol enzyme. (b) Set-up I: in solution, NADPH added. (c) Set-up II: immobilized P450 BM3 variants on ITO, NADPH added. (d) Set-up III: immobilized P450 BM3 variants on ITO, electric potential of −0.6 V applied. Reproduced from, ChemCatChem, 2018, 10, 525–530, with permission.234 Copyrights, 2018, Wiley.

With polyelectrolytes, LbL films can provide a partitioned microenvironment that is hydrated and conductive to allow fast diffusion of ions and substrates as discussed in Section 3.1. LbL films of bacteriorhodopsin with polycations had a high degree of orientation.235 Bacteriorhodopsin could be oriented by electrostatically adsorption from solution onto an electrode. The nonlinear optical coefficient from the second harmonic generation suggested that the second-order susceptibility of the LbL film was greater than that of an electric-field-oriented film. However, this is a special case, and enzymes in LbL films must rely on electron hopping mechanisms involving enzyme throughout the film as described above for thicker films but starting with electron transfer between electrode and the nearest enzyme to the electrode layer. All enzymes in the film may not be accessible for these reactions; in some cases, they may be too far away or improperly oriented from their neighbors. For example, in the LbL films of anionic glucose oxidase with anionic poly(allylamine) modified by ferrocene (PAA-Fc) on an Au electrode, a good fraction of glucose oxidase (∼40%) was inactive.236 To resolve this issue in LbL films, biocatalytic activity has been improved by the use of conductive nanomaterials, as discussed in Section 3.2. The formation of electrically conductive nanomaterial networks and the use of conductive polymers in LBL films237 can “wire” the electron transfer throughout the film. For example, 4 nm underlayers of conductive sulfonated polyaniline (SPAN) on rough PG were coated with LBL films of polyions and proteins HRP or Mb. With a SPAN underlayer present, >90% of protein electron transfer was coupled to the electrode, but only 40% of enzymatic electroactivity was found without SPAN.237 This allows utilization of nearly all the enzymes in the LbL film. In this way, the heterogeneous electron transfer rate constant (ks as discussed in Section 2.1) and the catalytic efficiency of enzymes (kcat, see Table 3) in these films can reach much higher values. However, these more complex approaches are not always necessary, since 40–50% of enzymes as active in LBL films are often sufficient for good catalytic performance in synthetic applications.

Table 3 Electrochemical transformations catalyzed by CYPs
Electrode Substrate Product Product identification methods K M k cat Ref.
CYP1A1/PBA-NGO/GCE Benzo[a]pyrene Benzo[a]pyrene-7,8-diol HPLC, GC/MS 26 μM 1.9 s−1 228
CYP2C9, CYP2C9*1, CYP2C9*2/4-aminothiophenol/Au electrode Tolbutamide 4-Hydroxytolbutamide HPLC 275 μM, 245 μM, and 722 μM 4.5 min−1, 2.3 min−1, and 1.9 min−1 274
CYP101/pyrene maleimide/GCE Camphor 5-exo-Hydroxycamphor Electrospray ionization mass spectrometry (ESI-MS) 275
CYP101/N-(4-carboxyphenyl) maleimide(p-CPMI) SWCNT/CGE Camphor 5-exo-Hydroxycamphor ESI-MS 190
CYP2C9/ITO/CS/GCE Tolbutamide 4-Hydroxytolbutamide HPLC, MS 203 μM 194
CYP2C9/AuNPs/TiO2 NTA Tolbutamide 4-Hydroxytolbutamide HPLC, MS 4.8 μM 9.89 s−1 176
CYP1A2 and CYP3A4/AuNP/CS/RGO Clopidogrel 2-oxo-clopidogrel by CYP1A2, and clopidogrel carboxylic acid by CYP3A4 LC-MS/MS 10.82 μM 2.42 s−1 122
CYP2C19/CS-ceria-graphene Omeprazole 5-Hydroxyomeprazole LC-MS, ESI-MS 8.22 μM 5.72 s−1 276
CYP2C9/DDAB/GCE S-warfarin 7-Hydroxywarfarin HPLC 4.28 min−1 277
Cyt P450 Vdh Vitamin D3 25-Hydroxyvitamin D3 HPLC 278
HLMs/PDA/Au@RGO Testosterone 6β-OH-testosterone HPLC-MS 175 μM 8.9 s−1 279
CYP2D6-CPR-PEI-RGO/GCE Tramadol O-Dimethyl-tramadol HPLC-MS 23.9 μM 0.47 s−1 230
CYP2D6/AuNPs-CS-RGO and CYP1A1/PAA-AuNPs-CS-RGO Tramadol O-Dimethyl tramadol LC-MS, GC-MS 280
PSS-(CYP1A2-PSS)2 Styrene Styrene oxide GC 226
PGE/MWCNT/HLM and PGE/MWCNT Testosterone 6β-OH-testosterone HPLC 290 ± 33 μM for PGE/MWCNT/HLM and 480 ± 51 μM PGE/HLM 281


4. Reaction conditions: solvent and beyond

Water is needed to hydrate enzymes and it is a major component in bioelectrocatalytic systems. Other components in the fluid may significantly influence the stability of the intermediates, reaction selectivity and yields of products.162 Supporting electrolyte in the solvents provides conductivity; however, it does much more than that. In general, any salts that dissociate to their ionic forms can work as a supporting electrolyte. The more frequently used supporting electrolytes contain cations including's K+, Na+, Li+, or H+ and anions ClO4, Cl, NO3, HPO42− or SO42−. The concentration of electrolytes is usually in the range of 0.1 to 1 M and they should be electrochemically inactive.133 Tuning the supporting electrolyte is important for maximizing the performance of biocatalysts. Specific ion effects are ions interacting specifically with enzymes, and should be avoided.133 While a higher concentration of electrolyte provides better conductivity, it may enhance background signals.

The use of organic solvents is sometimes needed because the solubility of substrates is limited in aqueous solution. However, enzyme stability issues typically rule out more than 10–20% of organic solvent in water; and at even these levels, organic solvents can denature enzymes by disrupting their internal hydrophobic interactions. Thus, organic solvents can be detrimental to the activity of enzymes. Easterbrook et al. demonstrated the influence of dimethyl sulfoxide (DMSO), acetonitrile and methanol (MeOH) in human hepatocytes on activities of Cyt P450s where organic solvents were present in water at of 0.1, 1 and 2% (v/v).238 DMSO showed the highest enzyme inhibitory effects for several Cyt P450s and activity retentions were approximately 40% (CYP2C9), 23% (CYP2C19), and 11% (CYP2E1) compared to those observed for 0.1% acetonitrile. MeOH showed a solvent-dependent drop of activity of CYP2E1 and CYP2C9, but it had less impact on other Cyt 450s, while acetonitrile in the range of 0.1–2 vol% had no observable effect on Cyt P450 activities.

Electron paramagnetic resonance (EPR) spectroscopy and Soret band optical absorbance suggested that the exposure of the organic solvents decreased the local polarity in enzyme active sites and shifted the Soret absorption band of the FeIIIheme center.239,240 The Soret band of horseradish peroxidase (HRP) in pH 7 buffer is 403 nm, but different microenvironments show a shift in the Soret absorption and intensity.241 The dissociation of FeIIIheme from enzymes leads to a blue shift to 398 nm. With organic solvents, the Soret band can increase in the intensity without shifting λmax.239 For example, in H2O/dioxane and H2O/methanol mixtures, an enhanced Soret absorbance at 403 nm was seen without a shift of λmax.239 However, in H2O/acetonitrile, a peak broadening was found in the range 370–425 nm. The native g values of EPR spectra of HRP in aqueous buffer (pH 7), are observed near g = 6 and g = 2 because of high spin ferric heme iron, with a rhombic splitting near g = 6.242 There was no change observed in the spectrum in 80% v/v dioxane, while with further increasing dioxane, the EPR spectrum broadened, suggesting the reduced spin-label flexibility. Note that extremely high amounts of organic solvents (e.g., 90% v/v acetonitrile, 95% v/v dioxane and 2% ethyl acetate and 1% v/v aqueous buffer) will directly change the enzyme structure and activity.

Hammett analysis was applied to examine the influence of solvent on the transition-state structure of HRP-catalyzed phenol oxidation.239 In this analysis, Hammett coefficients (ρ values) showed charge distribution in the reaction transition state, which is highly sensitive to both microenvironment of the transition state and the reaction mechanism. Therefore, variations in ρ value for HRP catalysis in organic solvents may suggest solvent penetration into the enzyme's active site.239 A deviation in Hammett ρ values reflect changes in the transition state. The catalytic activity of the enzymes decreased as much as 4 orders of magnitude in a mixed solvent containing organic solvents like dimethyl sulfoxide, dioxan, dimethylformamide, tetrahydrofuran and acetonitrile. The catalytic activity of HRP decreased by 30% with only 30 vol% organic solvents other than DMSO.243 In the case of DMSO, the activity of free enzyme and immobilized enzyme decreased with respect to concentration of DMSO. The stability of the enzyme can be improved by covalent immobilization on silica nanoparticles (NPs), where half-life in buffer at 50 °C was improved from 2 h to 52 h. These covalently immobilized enzymes demonstrated high catalytic activity and stability in comparison to free enzyme.243 The optimum activity of rice peroxidase on silica NPs was found at 30 °C at pH 7–8 and was stable up to 68 °C. Shelf-life of rice peroxidase was 60 h in 1,4-dioxane (20%) and 12 h in ethanol. The activity of the rice peroxidase was improved up to 12 h (shelf-life) with 0–40% ethanol or 0–30% 1,4-dioxane.244

Cyt 450 BM3 can catalyze a broad range of industrially significant compounds in water, but was rapidly inactivated in organic solvents.245 The impact of acetonitrile, ethanol, acetone, methanol and DMSO on the activity of Cyt P450s in rainbow trout hepatic microsomes was investigated.246 Acetonitrile, acetone, methanol, ethanol and DMSO below 0.5% did not affect the Cyt P450s BM3 activities. Methanol could be used up to 1% for CYP1A and CYP3A, and up to 2% acetone could be used with CYP2E1.

Cyt 450 BM3 can catalyze a broad range of industrially significant compounds in water, but was rapidly inactivated in organic solvents.245 The impact of acetonitrile, ethanol, acetone, methanol and DMSO on the activity of Cyt P450s in rainbow trout hepatic microsomes was investigated.246 Acetonitrile, acetone, methanol, ethanol and DMSO below 0.5% did not affect the Cyt P450s BM3 activities. Methanol could be used up to 1% for CYP1A and CYP3A,;and up to 2% acetone could be used with CYP2E1.

The pH of the reaction mixture is critical for maximizing activity and electrochemical behavior of the enzyme. Enzymes have a narrow pH range of activity, often near physiological pH and they can denature at temperatures above 40 °C. There are 7 isoenzymes of HRP and in solution they exhibit the maximum catalytic activity at pH 6.0–6.5.247 Enzymes in thin films may have slightly altered pH-activity profiles and redox potentials than in solution. Electrochemistry of bacterial Cyt P450cin (CYP176A), which hydroxylates cineol in DDAB films gave a significant pH dependence consistent with coupled electron/proton transfer at pH 6 to 10.248 FAD-reductase and P450BM3 immobilized in DDAB film electrodes had mid-point potential of −0.402 V and −0.244 V vs. SCE, respectively. A negative shift in the mid-point potential was found in 2 linear relations, one between pH 3 to 8 and another between 8 to 10 pH.249 In another study, CYP2C9 was immobilized in DDAB films on pyrolytic graphite electrodes. Ionic strength and pH affected the electrochemical behavior of the CYP2C9.250 The changes in pH led to a shift in midpoint potential of about −55 mV per pH unit, close to the theoretical value −59 mV per pH unit consistent with a single electron, single proton transfer process. In 0.1 M buffer, the change in pH led to a shift in midpoint potential of about −52.4 mV per pH unit (Fig. 8) between pH 5.8 and 8.0 in the absence and presence of substrate. Changes in ionic strength from 0.1 to 1 M showed slightly anodic shifted midpoint potentials, at pH 5.8, ∼22 mV and at pH 8.0, ∼9 mV. The electron transfer rate constant increased slightly with pH, 139 s−1 at pH 6.0, 150 s−1 at pH 7.4 and 165 s−1 at pH 8.5.


image file: d3cs00461a-f8.tif
Fig. 8 The effect of ionic strength and pH on the midpoint potential of purified CYP2C9 on DDAB modified PGE. The slope of each plot is 57.5 mV pH−1 for 1.0 M (open circles) and 52.4 mV pH−1 for 0.1 M (filled circles). Reproduced from Biochem. Pharmacol., 2005, 69, 1533–1541 with permission.250 Copyright, 2005, Elsevier.

To this end, microemulsions provide another possible solvent for bioelectrocatalysis, especially valuable for water-insoluble reactants. Microemulsions are clear, stable fluids made of oil, water, surfactant, cosurfactant and electrolyte featuring separate oil and water microphases.251,252 They can be water-rich, typically 30–80%, which provides hydration of enzymes, and an organic fluid to dissolve non-polar reactants. These fluids have the capability to solubilize polar, nonpolar, and ionic reactants. Cosurfactant is generally required to lower the interfacial tension essential for their formation. Water as the continuous phase contains salt or buffers that are required to enable conductivity necessary for electrochemical synthetic reactors and maintain high activity of enzymes. Structures feature oil droplets in water known as oil-in-water microemulsions or bicontinuous structures of tortuously intertwined mixtures of water and oil, which is often just a common organic solvent. Surfactants reside at oil–water interfaces and decrease interfacial free energy to near zero to stabilize the microemulsions, and additionally provide ionic conductivity for electrolyte. Controlling the compositions of these fluids can provide unique control over some reaction pathways in electrochemical synthesis.253–258 They are easily prepared from published phase diagrams by mixing components in the correct proportions.259,260

Our group pioneered heme enzyme bioelectrocatalysis in microemulsions.261 Microemulsions are stable mixtures of surfactant, co-surfactant, organic solvent, and water. Oil-in-water and bicontinuous microemulsions look like homogeneous solvents, but have internal microstructures or nanostructures of oil and water phases stabilized by surfactant layers at their interfaces. They are excellent solvents for nonpolar organic molecules, and provide water to hydrate enzymes in films. The use of microemulsions can increase enzyme reactivity, e.g., a 50-fold increase in the yield of styrene oxide from styrene epoxidation catalyzed by and enzyme-like reaction of FeIIIheme protein myoglobin (Mb) in CTAB/water/1-pentanol/tetradecane (17.5/35/35/12.5, wt%) microemulsions compared to the reaction in pH 7.4 buffer.261 The electrochemical reduction of trichloroacetic acid and 1,2-dibromocyclohexane catalyzed by Mb was also demonstrated in microemulsions.262 The binding and catalytic rates of Mb were more effective for reduction of 1,2-dibromocyclohexane in microemulsion (2000-fold) than in the buffer. For trichloroacetic acid, the faster rate was found in buffer solution (100-fold) because of its hydrophilicity and strong hydration in aqueous phase. The hydrophobic nature of 1,2-dibromocyclohexane provided better access to hydrated Mb in the surfactant film on the electrode and gave a faster reaction rate, while the hydrophilic trichloroacetic acid almost exclusively in the water phase had the poor access to catalyst Mb in the film. Cross-linked films of enzyme-polylysine (PLL) on electrode surfaces improved the thermal, and biocatalytic activity up to 90 °C.263 The cross-linked Mb/PLL film was stable in microemulsions, which improved the turnover rates for styrene epoxidation compared to loosely cross-linked or non-crosslinked Mb/PLL films.255 UV circular dichroism supported the stability and the stable conformation of the protein in the film at moderate to high temperatures. UV circular dichroism (CD) spectra of enzyme-PLL films (HRP/PLL, Mb/PLL, soybean peroxidase (SBP)/PLL) provided evidence of native protein secondary structure in the films.264 Circular dichroism spectral characteristics for the enzyme-PLL films appear at 193 nm maxima, and 210 and 222 nm minima, when in contact with microemulsion and buffer. All spectra suggested proteins retained native conformations in films (HRP/PLL, Mb/PLL, SBP/PLL) up to 90 °C in SDS microemulsion, and buffer. However, a small change in ratio of 210/222 nm minima was found with CTAB microemulsion in the HRP/PLL films (Fig. 9).


image file: d3cs00461a-f9.tif
Fig. 9 UV CD spectra of enzyme-PLL films on fused silica slides: (a) HRP/PLL, and (b) Mb/PLL films in SDS, and (c) HRP/PLL, and (d) SBP/PLL films in CTAB microemulsions at 25 and 90 °C. UV CD spectra recorded with pH 5.5 buffer + 0.1 M NaCl at 90 °C were added for comparison. Reproduced from, Langmuir, 2008, 24, 10365–10370, with permission.264 Copyrights, 2008, ACS Publications.

We used tert-butylhydroperoxide (t-BuOOH) to examine the catalytically active ferryloxy Mb radical formation.265 Mb/PLL films on a PG electrode were reacted with t-BuOOH to generate ferryloxy Mb, and the reaction followed Michaelis–Menten enzyme kinetics. Heterolytic peroxide cleavage then formed tert-butyl alcohol and Mb ferryloxy radical, while homolytic cleavage led to formation of the alkoxy radical and nonradical MbFe(IV)[double bond, length as m-dash]O. The homolytic/heterolytic product ratio was ∼0.50, and ∼67% products were produced by the heterolytic pathway. Catalytic current from RDV supported the generation of ferryloxy radical, and peroxide-initiated styrene epoxidation in the presence of the t-BuOOH by the Mb/PLL films. RDV studies were used to measure catalytic rates of Mb catalyst in CTAB and SDS microemulsions and buffers. Acidic buffer and acidity of the SDS microemulsions had an effect on the reaction kinetics. The values of kcat and KM were larger in neutral water phases. Values of kcat were close to each other for fluids with pH 2 and 12 water phases. Buffer at pH 6.5, neutral water SDS microemulsion (W/O), and neutral water CTAB micelles, had the largest kcat values.

5. Reactions catalyzed by Cyt P450s

Organic molecules that are commonly unreactive with most chemical oxidants can be activated by Cyt P450s (Scheme 2).266,267 High-valent iron-oxo species in these enzymes are strong oxidizing agents capable of oxidizing inert C–H bonds through hydrogen atom abstraction, e.g., transforming poorly reactive linear or cyclic alkanes to alcohols and alkenes to epoxides.267–273Fig. 10 summarizes bioelectrochemical transformations catalyzed by Cyt P450s, including heteroatom oxygenation, C-hydroxylation, S-oxidation, N-hydroxylation, epoxide formation and heteroatom dealkylation. Additionally, Cyt P450s have also been used in biosensors to produce drug metabolites and inhibitors and some of those representative compounds are also reviewed in this section. The successful reactions catalyzed by Cyt P450s on electrode surfaces are listed in Table 3. The Cyt P450 electrodes used for biocatalytic reactions with KM, kcat, substrate and respective products are also summarized in Table 3.
image file: d3cs00461a-f10.tif
Fig. 10 Schematic representation of products of bioelectrocatalytic reactions catalyzed by Cyt P450s. The down arrow near the electrode with H2O2 on it represents the H2O2 shunt. With Cyt P450s in microsomes or supersomes, electron transfer proceeds first to CRP, and CPR donates electrons to the Cyt P450 iron heme as shown in Scheme 2 and discussed previously. The same products can form by either pathway.

5.1 Activation of C–H bonds

Cyt P450s metabolize drugs and xenobiotics (Scheme 1) in mammals through C–H bond activation.282–286 Hydroxylation of C–H on organic reactants is one of the most common reactions catalyzed by Cyt P450s with addition of an OH group that can be regio- and stereo-selective. The stereoselectivity of enzymatic catalysis usually arises from specific binding affinities of substrates with different chirality. In the case of most natural Cyt P450s, the heme center is more selective for binding S*-isomers than R-isomers, and/or to produce predominantly an S*-isomer product from an achiral reactant, although this might vary for different species of Cyt P450s. CYP2C9, as an example, can catalyze the hydroxylation of 5,5-diphenylhydantoin (PPT) to produce the 5-(4′-hydroxyphenyl)-5-phenylhydantoin (p-HPPH) through an arene-oxide insertion. CYP2C9 is ∼40 times more stereoselective towards S-p-HPPH,287 but the S/R isomer ratio is dependent on the genetic polymorphism of CYP2C9. In case of oxidation of nicotine by CYP2A6, computational simulation suggests hydrogen transfer from either the trans-5′- or cis-5′-position of (S)-(−)-nicotine.288 The hydroxylation process involves two reactions, starting with the hydrogen transfer from the 5′-position of (S)-(−)-nicotine to the oxygen of compound I (known as the H-transfer step), followed by recombination of the (S)-(−)-nicotine moiety with an iron-bound hydroxyl group, resulting in the formation of the 5′-hydroxynicotine product (known as the O-rebound step). During the 5′-hydroxylation process, there is a preferential loss of the trans-5′-hydrogen in a stereoselective manner, because the trans-5′-hydrogen is spatially closer to the oxygen of Cpd I, as compared to the cis-5′-hydrogen. The calculated product stereoselectivity of S/R isomers for the overall process is 97%, closely matching with the experimentally observed stereoselectivity of 89–94%.

Biologically relevant molecules, e.g., drugs, usually go through metabolic reactions catalyzed by Cyt P450 to form non-toxic or toxic metabolites. This process can play a significant role in development of safe and high-value drugs.289,290 CYP2C9, as an example, can catalyze the metabolic oxidation of tolbutamide.274 CYP2C9 was assembled on a p-aminothiophenol-modified Au electrode. Thiol derivatives like 4-aminothiophenol, thiophenol, 4-hydroxythiophenol and 4-carboxythiophenol were attached by Au-thiolate bonds overnight in 0.1 mM solutions in ethanol. CYP2C9 microsomes were drop-cast onto those thiol-modified Au electrodes and incubated for 10–30 min at 25 °C. Well-defined redox responses were only found on electrodes modified by thiophenol or 4-aminothiophenol and then coated with CYP2C9 at pH 7.3. The peak potential and the charge transfer rate constant (ks) were −0.399 V vs. Ag/AgCl and 300 s−1 for CYP2C9/4-aminothiophenol/Au electrode; and −0.330 V vs. Ag/AgCl and 200 s−1, at CYP2C9/thiophenol/Au electrode, respectively. The oxidation of tolbutamide was done at −0.45 V vs. Ag/AgCl for 2 h with CYP2C9 with 500 μM tolbutamide in pH 7.3 buffer containing 0.5 vol% of ethanol to give product 4-hydroxytolbutamide. The kcat and KM for CYP2C9 were 4.5 min−1 and 275 μM for CYP2C9*1; 2.3 min−1 and 245 μM for CYP2C9*2; and 1.9 min−1 and 722 μM for CYP2C9*3, respectively. Similarly, CYP2C9 with CRP coated on ITO with chitosan (CS)/GCE was used to oxidize tolbutamide to 4-hydroxytolbutamide (Fig. 11).194 The oxidation was studied by titration of tolbutamide at −0.48 V in the buffer solution. There was a clear increase of the cathodic current that reached a stable value within less than 5 s (Fig. 11A). The current increase was due to the catalytic conversion of tolbutamide. The current vs. concentration curve was fitted with the Michaelis–Menten model to obtain a Michaelis constant (KM, characteristics to analyze the binding of substrates to enzymes) of 203 μM (Fig. 11B), close to the KM of CYP2C9 (60 to 400 μM) in human liver microsomes. The electrolysis was done at −0.48 V vs. SCE with 600 μM tolbutamide to give 4-hydroxytolbutamide.


image file: d3cs00461a-f11.tif
Fig. 11 Bioelectrocatalysis of tolbutamide oxidation: (A) Current vs. time plot of consecutive addition of tolbutamide in 0.1 M PBS (pH 7.4) with saturated O2. (B) Calibration plots demonstrating current vs. concentration of tolbutamide. Reprinted from Chem. Commun., 2012, 48 (63), 7802 with permission.194 Copyright, 2012, Royal Society of Chemistry.

Other approaches have been developed using enzymes with nanomaterials on surfaces of electrodes to improve turnover numbers, and selectivity. When enzymes are immobilized on electrodes, their electrochemical activity depends upon various factors discussed in Section 3. The electrodes with different nanomaterials, nanostructured materials, or ionic surfactants including carbons, conductive polymers, polyions, DDAB, etc., generally improved the activity of Cyt P450s as discussed in Section 3.226,228,291 The crystallinity of electrode materials varies the activity as well, e.g., a higher electroactivity of CYP3A4 was fund on polycrystalline indium tin oxide (ITO) as compared to that supported on amorphous ITO.221 CYP3A4 immobilized on polycrystalline ITO enhanced the rate of oxygen reduction and specific coupled with oxidation of drugs (quinidine and testosterone) by oxidation. Similarly, CYP3A4 immobilized with DDAB on a graphite screen-printed electrode (SPE) was highly active for hydroxylation of diclofenac.292 CYP1A1 on the pyrenebutyric acid modified nitrogen-doped graphene composite (PB-NGO) electrodes can catalyze the oxidation of benzo[a]pyrene (B[a]P) to benzo[a]pyrene-7,8-diol.228 The redox peak was seen for CYP1A1/PB-NGO/GC electrode at a potential of −0.48 V, while the peak to peak separation was 56 mV at a scan rate of 100 mV s−1 in 0.1 M phosphate buffer (pH 7.4) under nitrogen to confirm a one-electron transfer. The Michaelis constant (KM) was 26 μM and the catalytic rate constant (kcat) was 1.9 s−1. In addition, the linkage of enzymes to electrode can also change the enzymatic activity. For example, bacterial Cyt P450 (CYP101) immobilized on GC through pyrene maleimide enhanced the activity for hydroxylation of camphor.275 The rate constant of electron transfer (ks) was ∼2.5 times higher, in comparison to non-specifically bound enzymes.

Nanostructured materials can usually provide a large surface area to improve the activity of Cyt P450s, e.g., on the surface of nanostructured Au and reduced graphene oxide. CYP2C9 when immobilized on AuNPs supported on TiO2 nanotube arrays (NTAs) was highly active for metabolic reactions of tolbutamide (Fig. 12).176 NTAs were synthesize by anodization where hollow TiO2 nanotubes with diameters of 50–100 nm and length up to a few micrometers could be prepared on a Ti foil. Afterwards, the electrochemical deposition of AuNPs (∼5.5 nm) on TNAs was performed by cyclic voltammetry. CYP2C9 was immobilized on AuNPs/TNAs/Ti foil by dip coating in the solution of 1.0 μM CYP2C9 in phosphate buffer 7.4 at 4 °C for 24 h. The electron transfer rate constant (ks) of CYP2C9 was calculated to be 7.8 s−1 at a scan rate of 0.1 V s−1, about 1.3 times greater than CYP2C9 coated on an Au electrode directly. The Michaelis constant (KM) increased with the tube aspect ratio where the binding of substrates become more difficult in NTAs with a larger diameter. As compared to CYP2C9 directly immobilized on the surface of Ti foil by electrostatic interaction, the KM decreased by 10-folds from 145.6 μM to approximately 10 μM, indicating that the high surface area favors the binding of substrates. The kcat could reach 9.9 s−1 at an optimized tube aspect ratio to produce 4-hydroxytolbutamide. Other conductive nanomaterials e.g., reduced graphene oxide can be used to improve the activity of enzymes as well. CYP2C19 can be immobilized on reduced graphene oxide and CeO2 nanocomposites for metabolic oxidation of omeprazole into 5-hydroxyomeprazole.276 The electron transfer rate constant (ks) of CYP2C19 was 1.77 s−1 at the scan rate of 0.1 V s−1. For the Michaelis–Menten fitting of catalytic current and reactant concentration, the Michaelis constant (KM) of omeprazole to CYP2C19 was calculated to be 8.22 μM and the catalytic rate constant kcat was 5.72 s−1, indicating the faster turnover of substrates. The yield of 5-hydroxyomeprazole was 17% after 4 h electrolysis at −0.520 V vs. SCE.


image file: d3cs00461a-f12.tif
Fig. 12 Synthesis of TiO2 nanotube arrays system for investigation metabolite of tolbutamide into 4-hydroxytolbutamide. Regenerated from Anal. Chem., 2014, 86(15), 8003–8009 with permission.176 Copyright, 2014, ACS Publications.

Cyt P450s are active for oxidation of natural drugs as well. Vitamin D3 (VD3) is essential for calcium metabolism in the human body. It is not active in its native form and the activation through 25-hydroxylation catalyzed by Cyt P450s is needed for its biological activity. Human CYP3A4, CYP3A5 and CYP3A7 as examples are active to convert VD3 to 25-dihydroxyvitamin D3 (or denoted as 25(OH) D3).96 The bioelectrocatalysis method was used for this hydroxylation of VD3 into 25(OH) D3 by using ferredoxin and other molecules as mediators with Cyt P450 vitamin D3 hydroxylase (Cyt P450 Vdh) (Fig. 13).278 In solution, Cyt P450 VdH with NADPH produced 2.3 nmol of 25(OH) D3 after 30 min using NADH as an electron source. However, Cyt P450 VdH produced 4.8 nM of 25(OH) D3 at a 0.9 cm2 electrode by electrolysis at −0.5 V vs. SHE for 30 min. The yield of 25(OH) D3 was nearly doubled in later case without the use of NADPH.


image file: d3cs00461a-f13.tif
Fig. 13 Conversion of vitamin D3 to 25-dihydroxyvitamin D3 by Cyt P450 vitamin D3 hydroxylase.

Cyt P450s in human liver microsomes (HLMs) can catalyze the conversion of testosterone into 6β-OH-testosterone on a polydopamine decorated Au-graphene (PDA/Au@RGO) nanocomposite electrode.279Fig. 14A shows electrochemical behavior of the HLMs on PDA/Au@RGO. The electrode showed the well-defined redox peak with an anodic peak at −0.37 V and a cathodic peak at −0.45 V vs. SCE, under anaerobic conditions, while no peak was observed without microsomes. Fig. 14B shows that the enzymatic electrocatalytic oxygen reduction current and binding of oxygen to reduced iron heme was observed at −0.5 V under aerobic conditions (curve b). However, the oxygen reduction currents on PDA/Au@RGO-CS surface with microsomes were very low. Thus, microsomal heme proteins catalyzed the oxygen reduction through electron transfer from electrode to reduce CPR then to the heme. In curve c (Fig. 14B), the cathodic current increased on addition of testosterone into the electrolyte, known as a catalytic process. In control study, the testosterone did not make any significant changes in the current (curve c in the insert of Fig. 14B). Electrolysis at −0.5 V in oxygen saturated phosphate buffer was done for testosterone (250 mM) in water/methanol (0.5%) for 2 h. The rate constant and KM were 8.9 s−1 and 175 μM, respectively. The product was confirmed by HPLC-MS (Fig. 14C) and MS (Fig. 14D) where 305.4 (m/z) and 343.0 (m/z) were assigned to [M + H]+ and [M + K]+ of 6β-OH-testosterone, respectively.


image file: d3cs00461a-f14.tif
Fig. 14 (A) CVs on (a) without microsomal PDA/Au@RGO-CS/GCE, (b) with microsomal PDA/Au@RGO-CS/GCE and (c) background subtracted response of curve b in anaerobic 0.1 M PBS (pH 7.4). (B) Cyclic voltammograms on microsomal PDA/Au@RGO-CS/GCE, (a) anaerobic, (b) aerobic PBS and (c) with testosterone (200 mM) Insert: curves on without microsomal PDA/Au@RGO-CS/GCE in (a) anaerobic, (b) aerobic PBS and (c) with testosterone (200 mM). (C) HPLC of the electrolysis at −0.5 V for 2 h and (D) the mass spectra of the product observed after the reaction. Regenerated from Electrochimica Acta, 2017, 258, 1365–1374 with the permission.279 Copyright, 2017, Elsevier.

As mentioned previously, our group used LbL films of DNA, an ECL-generating polymer [Ru(bpy)2(PVP)10]2+ {(PVP = poly(4-vinylpyridine)} (Ru-PVP) and many different human CYPs in supersomes contain CPR on multi-well microfluidic arrays built on top or PG electrodes to make metabolites that damage DNA. Test compounds flow into the array, and the bioelectrocatalytic enzyme film under potential control produces the metabolites close to the DNA layers. If the metabolite is reactive, DNA is damaged and detected by a subsequent electrochemiluminescence measurement in the same array as described earlier.23–25 Multiple wells in the array can accommodate many different types of Cyt P450s in a single experiment (Fig. 15).293 Here we provide a few examples of this approach. A microfluidic 64-nanowell chip array was used to evaluate genotoxicity of metabolites formed by liver, lung, kidney and intestinal Cyt P450 enzymes. LbL films contained DNA, RuBPY and the metabolic enzymes. Multiple enzyme reactions for a reactant are run via bioelectrocatalysis, then ECL measuring the DNA damage is recorded with a sensitive camera in a dark box. LC-MS studies of the same reactions are used for confirmation of array results, and to detect specific products. Results reveal rates and nucleobase adducts from DNA damage, enzymes responsible from different organs, and pathways of genotoxic chemistry.293 DNA damage rates were found using the cell-free array and correlated with organ-specific cell-based DNA damage. In DNA, deoxyguanosine is the oxidized into 8-oxo-7,8-dihydroxy-2-deoxyguanosine (8-oxodG) by reactive oxygen species, which is an important oxidative DNA damage product in humans.294,295


image file: d3cs00461a-f15.tif
Fig. 15 Conceptual representation of (a) microfluidic bioelectrocatalytic arrays to determine DNA damage by ECL, combined with (b) biocatalytic reactive metabolite-DNA adduct formation and measurements by LC-MS/MS; (c) representation of reactive metabolite-DNA adducts formation; (d) ECL chip contains a flow cell, and 64-well PG electrode with each well containing LbL films of different CYPs, DNA, and Ru-PVP. Reproduced from Chem. Sci., 2015, 6, 2457–2468 with permission.293 Copyright, 2015, RSC.

We developed similar arrays to measure DNA oxidation facilitated by redox-active metabolites formed by the Cyt P450s. The metabolites of nitrosamines from cigarette smoke can lead to DNA oxidation and genotoxicity.296 Metabolites of nitrosamines 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone (NNK) and N-nitrosonornicotine (NNN) were formed by bioelectrocatalysis in a 3d-printed microfluidic array loaded with thin films of DNA, CYPs and other enzymes that make metabolites in presence of Cu2+ and NADPH to obtain relative rates of DNA oxidation. Here, electrochemiluminescence (ECL) measurements employing a soluble ECL dye, [Os(tpy-benz-COOH)2]2+ were carried out to detect the primary DNA oxidation product 8-oxo-7,8-dihydro-2-deoxyguanosine (8-oxodG, Fig. 15). Results showed that metabolites of NNK and NNN + Cu2+ + NADPH generated high rates of DNA oxidation by forming reactive oxygen species (ROS) in a complex pathway as the ultimate oxidants.296

In biological catalysis, multiple enzymes can mediate multi-step conversions. Clopidogrel as a thienopyridine antiplatelet agent is used for treatment of cardiovascular diseases. The activation of clopidogrel is a two-step reaction both catalyzed by Cyt P450s.297 The first oxidation is clopidogrel to 2-oxo-clopidogrel as a sp2 C–H activation and the second step is the ring opening of 2-oxo-clopidogrel into a thiol-containing active form (Fig. 16). Metabolic reaction of clopidogrel was done using a bi-enzymatic system on the surface of a AuNP/chitosan/reduced graphene oxide (RGO) nanocomposite (Fig. 16).122 Two enzymes, namely CYP1A2 and CYP3A4, were co-immobilized on an electrode to catalyze the two cascade reactions. CYP1A2 was covalently attached to primary amine sites of chitosan with glutaraldehyde and CYP3A4 by hydroxyl sites of chitosan via amide bonds using carbonyldiimidazole. Enzymes were immobilized on the surface to provide consecutive oxidation of clopidogrel into 2-oxo-clopidogrel catalyzed by CYP1A2 and 2-oxo-clopidogrel converted into clopidogrel carboxylic acid catalyzed by CYP3A4. The co-loading of the two enzymes provides good electron communication to the two heme centers with a fast heterogeneous electron transfer rate constant (ks) 13.09 s−1. The Michaelis constant (KM) of clopidogrel metabolism by the two enzymes was calculated to be 19.6 μM and the catalytic rate constant kcat was 1.33 s−1, suggesting the two enzymes working cooperatively. The conversion of clopidogrel to clopidogrel carboxylic acid was up to 7.2% in 1 h electrolysis at −0.53 V (vs. SCE).


image file: d3cs00461a-f16.tif
Fig. 16 Electrochemical investigation of metabolites of clopidogrel by using 2 Cyt P450 enzymes at AuNPs/chitosan/rGO/GCE. Reproduced from Electrochimica Acta, 2015, 165, 36–44 with permission.122 Copyright, 2015, Elsevier.

In native Cyt P450s, high-valent iron-oxo species oxidize C–H bonds by oxygen transfer as discussed previously (Scheme 2). Conversion of C–H into C–OH bonds was been reported even for non-biologically relevant substrates, e.g., 4-cyclohexylbenzoic acid hydroxylation by CYP199A4,298 butane and propane hydroxylation by CYP450cam,299 methane oxidation by CYP153A6,300 ethane and octane hydroxylation by P450 BM3,301,302 cyclohexane hydroxylation by Cyt P450 from Acidovorax sp. CHX100 in recombinant P. taiwanensis VLB120,303 and Synechocystis sp. PCC 6803,304 and toluene oxidation by P450 2A13.305 These non-biologically relevant substrates may have weak binding to Cyt P450s as not being their biologically relevant function. Therefore, the yield is low in solution or in electrosynthesis.298,301 Similarly, the amination of C–H bonds by Cyt P450s is difficult with low selectivity.306 To increase the reactivity of enzymes, protein engineering on either heme or the secondary coordination sphere of heme has been used. Engineered enzymes can provide customized binding pockets for non-biologically relevant substrates. For example, replacing the heme cofactor of CYP119 by an iridium porphyrin (Ir(Me)-PIX), Ir(Me)-CYP119 can provide catalytic insertion of nitrenes into C–H bonds in sulfonyl azide derivatives with an enantiomeric excess of 95[thin space (1/6-em)]:[thin space (1/6-em)]5, er.307 Similarly, the modification of the secondary coordination sphere close to the Fe-heme center can enhance the demethylation activity of CYP2D6.10. CYP2D6-C, with valine (V) at position 122 instead of alanine (A) present in CYP2D6.10 was twice as active in converting codeine into morphine with a yield of >70%.308 CYP2C9 and its polymorphic variants CYP2C9-FLD prepared through CYP2C9 gene fusion at a genetic level with D. vulgaris flavodoxin (FLD) redox module showed altered catalytic activity to oxidize S-warfarin into S-7-hydroxywarfarin (Fig. 17).277 CYP2C9-FLD immobilized in DDAB films on a GC electrode (DDAB/GCE) was 4 times more active for this hydroxylation reaction, as compared to CYP2C9 under the same conditions.


image file: d3cs00461a-f17.tif
Fig. 17 Electroenzymatic conversion of S-warfarin into 7-hydroxywarfarin at DDAB/GCE.

5.2 Epoxidation

Epoxides are useful synthetic precursors to produce chiral molecules via ring-opening. Chiral epoxides are used widely in the pharma industry for synthesis of drugs.309,310 Manganese (Mn) chiral salen complexes are useful for enantioselective epoxidations.311 Epoxidation is considered to proceed through multiple mechanisms and different reactive intermediates.311,312 The enzyme can catalyze the oxygenation of alkenes (non-chiral substrates) to produce chiral epoxides.40,313–315 Predicting epoxide formation in drugs usage is difficult but important because the formation of epoxide is often linked with drug toxicity.316,317 Cyt P450s are capable of converting alkenes into epoxides.316,318–323 Several chemical reactions based on enzymes Cyt P450 BM3, CYP152A1, CYP199A and CYP199A4 have been reported for epoxidation, e.g., linolenic acid, linear aliphatic alkenes, 4-vinylbenzoic acid and styrene. Cyt P450 BM3 can catalyze stereo- and regio-selective epoxidation of linolenic acid to epoxy-linolenic acid.324 The epoxidation is highly regioselective (100%) where only the C15[double bond, length as m-dash]C16 vinyl group is converted among the three vinyl groups in linolenic acid. The configuration of epoxy-linolenic acid is identified to be 15(R),16(S)-epoxyoctadeca-9,12-dienoic acid with an enantiomeric ratio excess of 60%. Epoxidation of 4-vinylbenzoic acid was reported using bacterial CYP199A4 as a biocatalyst with a high stereoselectivity and activity to produce (S)-epoxide with an enantioselectivity of 99%.318 In another approach, H2O2 also drives the epoxidation of styrene to styrene epoxide catalyzed by Cyt P450s, although selectivity is much lower.325 However, most alkenes are very hydrophobic with a limited solubility in aqueous reaction media. At a low concentration of substrates, Cyt P450s also reduce oxygen through two-electron pathways to produce H2O2 that can oxidize alkenes directly or activate the Cyt P450 via the peroxide shunt (Scheme 2) to epoxide or aldehyde as a mixture.309,325,326

In our group, the LBL films of pure CYP1A2, and bacterial Cyt P450cam with conductive polymer PSS on a carbon cloth electrode were used for electrochemical oxidation of styrene to styrene oxide.226 The epoxidation was facilitated at −0.6 V vs. SCE by initial biocatalytic reduction of molecular oxygen to H2O2 which activates iron heme enzymes to PFeIV[double bond, length as m-dash]O by a complex mechanism (Fig. 10) for catalytic oxidation of styrene. Here, iron porphyrin (PFeIII) reacts with bioelectrocatalytically-formed hydrogen peroxide to form high valent PFeIV[double bond, length as m-dash]O oxidant, which oxidizes styrene to styrene oxide. The turnover rate was 39 h−1 to produced styrene oxide at PSS-(CYP1A2-PSS)2 by electrolysis at −0.6 V vs. SCE for 1 h. The yield of styrene oxide and benzaldehyde were found to be 8.8 and 22 nmol, respectively.

5.3 Oxidative N-, and O-dealkylation

Products formed through O-dealkylation and N-dealkylation follow similar reaction mechanisms to remove alkyl groups to yield alcohols, carbonyls, or amines. Oxidative N-dealkylation is commonly catalyzed by Cyt P450s in secondary and tertiary amines, and related N-alkylated amines.327 In the first step, hydroxylation takes place at an α-carbon atom or carbon-adjacent heteroatom, leading to the formation of carbinol species which further undergo chemical rearrangement followed by dealkylation on the heteroatom. The dealkylation can be initiated either by single electron transfer or hydrogen atom transfer.328–335 For oxidative N-dealkylation,336–339 two elementary steps are involved, i.e., hydrogen atom transfer and single electron transfer. N-Oxidation proceeds through charge transfer to the oxo-ferryl group, followed by the formation of radical cation and homolysis of the iron–oxygen bond to generate the new N–O bond. For example, O-dealkylation of tramadol was done using CYP2D6 and CPR in a PEI film on graphene-modified GCE (Fig. 18).230 The electron transfer rate constant and Michaelis constant (KM) were 0.47 s−1 and 23.9 μM, respectively. Electrolysis was done at −0.52 V vs. SCE for 2 h at CYP2D6/PEI-RGO/GC electrode (RGO = reduced graphene oxide) with 300 μM tramadol. The conversion of tramadol into O-desmethyl tramadol was confirmed by MS with a m/z of 250.18 and the yield was ∼15% after electrolysis for 2 h.
image file: d3cs00461a-f18.tif
Fig. 18 Scheme of bioelectrocatalytic oxidation of tramadol. Reproduced from Langmuir, 2014, 30, 11833–11840, with permission.230 Copyright, 2014, ACS Publications.

Bioelectrochemical conversion of tramadol to O-desmethyl tramadol was done by using CYP2D6 and CYP1A1 immobilized on AuNPs-chitosan (CS), RGO, and polyacrylic acid on GCE (Fig. 19).280 The modification of AuNPs-CS-RGO/GCE with PAA gave the net negative charge on the electrode that further immobilized CYP2D6 and CYP1A1 through electrostatic interaction. PAA, while increasing the size of the gold nanoparticles from ∼15 to ∼25 nm, led to the potential shift from −492 to −513 mV for CYP2D6 and −504 mV to −509 mV for CYP1A1, respectively. The presence of PAA resulted in the decrease of the electron transfer rate constant (ks) from 5.19 s−1 to 4.10 s−1 for CYP2D6 and 3.24 s−1 to 2.78 s−1 for CYP1A1, respectively. The electrolysis converted tramadol to O-desmethyl tramadol at −0.530 V for 1 h, as confirmed by HPLC, GC-MS and LC-MS. The conversion per unit time was 2.3 h−1 for CYP2D6/PAA-AuNPs-CS-RGO and 3.2 h−1 for CYP2D6/AuNPs-CS-RGO.


image file: d3cs00461a-f19.tif
Fig. 19 Immobilization of enzymes on nanocomposite modified electrode and electrochemical system for metabolism of tramadol. Reproduced from J. Electroanal. Chem., 2016, 772, 46–51 with permission.280 Copyright, 2016, Elsevier.

6. Reactions catalyzed by peroxidases

As mentioned earlier, peroxidases are universal oxidoreductases that use hydrogen peroxide or organic hydroperoxides to oxidize prosthetic group metal atoms to higher valent oxidants. These enzymes catalyze reactions via free radicals where reactants get oxidized or in some cases polymerized (Fig. 20). Peroxidases are synthesized by bacteria, fungi, and plants. The reactive centers can be heme, selenium, cysteine thiols, manganese, and other chemical moieties.340 Several types of peroxidases have been used frequently for catalyzing organic and inorganic reactions such as Fe heme-containing (e.g., HRP, soybean peroxidase (SBP), chloroperoxidase (CPO)) and non-heme containing (e.g., manganese peroxidase, lignin peroxidase). Heme-containing peroxidases predominantly oxidize Fe(III) into highly reactive Fe(IV)[double bond, length as m-dash]O species by reducing hydrogen peroxide or other peroxides and catalyzing the reactions as discussed in Section 1 (see eqn (1)–(3)). However, the activation of peroxidases requires the use of H2O2 to produce the oxidative compound I form. In bioelectrochemistry, in situ H2O2 production through electrochemical oxygen reduction can also be used to activate the heme center for catalytic conversions. There is a fundamental need for direct electrochemical activation of peroxidases on an electrode for catalytic oxidation of organic compounds.341
image file: d3cs00461a-f20.tif
Fig. 20 Bioelectrocatalytic reactions catalyzed by peroxidases.

HRP is obtained from the roots of the horseradish plant,342 and has been widely used in biocatalytic oxidations of many organic and inorganic substrates.57 SBPs obtained from soybean plants carry out reactions similar to HRP.343 Chloroperoxidases are heme-containing enzymes found in fungi and catalyze chlorinations.344 In the same fashion, non-heme containing peroxidases, not our focus in the current review, also generate highly reactive species by reaction with peroxides. Manganese peroxidase as an example is a lignin-modified peroxidase found from wood-colonizing basidiomycetous fungi.345 This non-heme peroxidase is oxidized from Mn(II) to highly reactive Mn(III) by peroxides. It is stabilized by chelation with dicarboxylic acid and catalyzes the oxidation of phenolic lignin.

While pure peroxidases are less expensive compared to Cyt P450s, they are unstable under a higher temperature (e.g., >40 °C) or in the presence of excess peroxide.346,347 In our group, Guto demonstrated that chemical cross-linking in an LbL film can stabilize peroxidases, like HRP, Mb and SBP, for biocatalytic applications up to 90 °C.263 In the thin crosslinked LbL films with PLL on PG electrodes, HRP, SBP and Mb were catalytically stable for about 9 h at 90 °C; but were denatured completely in 20–30 min at 90 °C in solution. The HRP-PLL film was 3-fold more active than the SBP-PLL film at 25 °C; whereas, at 90 °C, SBP gave 8-fold better catalytic activity compared to HRP-PL. This is related to the inherent temperature stability of SBP as well as stabilization by the film where the unfolding of enzymes is limited by chemical cross-linking. The oxidation of o-methoxyphenol to 3,3′-dimethoxy-4,4′-biphenoquinone by crosslinked peroxidase-PLL films also showed better yields at 90 °C than 25 °C. Results demonstrated that peroxidase catalytic activity can be enhanced at elevated temperatures. Microemulsions of oil, water, and surfactant were used as supporting electrolytes for biocatalysis and bioelectrocatalysis at low and high temperature by using PLL-HRP, PLL-SBP and PLL-Mb activated by t-BuOOH.264 The enzyme-PLL films were stable at 90 °C for >12 h in microemulsions. Oxidation of o-methoxyphenol into 3,3′-dimethoxy-4,4′-biphenoquinone gave 3–5 times better yield in microemulsion at 90 °C than at 25 °C. Product yields increased 5-fold with increasing the temperature from 25 to 90 °C for both HRP and SBP in CTAB microemulsions and ∼3 fold in SDS microemulsions.

As mentioned, peroxides may be unstable and potentially inactivate by peroxidases at a higher concentration, although these enzymes are stable at mM peroxide concentrations. Continuous electrocatalytic synthesis of hydrogen peroxide by reduction of oxygen is a strategy to activate peroxidases while avoiding high levels of peroxide. Using dissolved oxygen in water, in situ reduction can generate hydrogen peroxide by, (i) a two-electron process at the electrode surface as a cathode, or (ii) enzymatic reduction of oxygen by glucose oxidase.348–354In situ generation of hydrogen peroxide was effective for activation of peroxidase enzymes. For example, in situ generated hydrogen peroxide was achieved at pH 3.0 at applied potentials +0.2 to −0.4 V vs. Ag/AgCl by oxygen reduction by using Pt-coated titanium and vitreous carbon.355 The generation of hydrogen peroxide increased at the more negative potential, but a higher concentration of hydrogen peroxide decreased the activity of catalyst lignin peroxidase. The optimal potential was +0.1 V vs. Ag/AgCl for generation of hydrogen peroxide and bioelectrocatalytic degradation of 2,4,6-trinitrotoluene by lignin peroxidase into 4-diamino-6-dinitrotoluene. The peroxidase-driven bioelectrocatalytic reactions are summarized in Table 4.

Table 4 Electrochemical transformations catalyzed by peroxidase
Electrode Substrate Product Product identification methods Yield Ref.
CS-DDAB-CPO-Nafion/GCE Cinnamyl alcohol Cinnamyl aldehyde GC-MS 52% 356
LPO Veratryl alcohol Veratraldehyde UV-vis 30 μM min−1 357
CPO Thioanisole (R)-Methylphenylsulfoxide GC, NMR 30 g L−1 d−1 341
CPO Thioanisole (R)-Methylphenylsulfoxide HPLC 104 g L−1 d−1 358
CPO Methyl p-tolyl sulfide (R)-Methyl p-tolyl sulfoxide GC, NMR 76% (ee 93%) 359
N-MOC-L-methionine methyl ester (R)-N-MOC-L-methionine methyl ester sulfoxide 60% (dr 81[thin space (1/6-em)]:[thin space (1/6-em)]19)
1-methoxy-4-(methylthio)benzene (R)-Methoxyphenylmethylsulfoxide 83% (ee 99%)
CPO Thioanisole Methyl phenyl sulfoxide GC-MS 23 g L−1 d−1 360
Monochlorodimedone Dichlorodimedone 9.5 g L−1 d−1
Indole Oxindole 8.3 g L−1 d−1
CPO-Au@HMS Ethylbenzene R-1-phenylethanol GC ∼1.6 mM in 12 h 361
AaeUPO Ethylbenzene 1-Phenethyl alcohol, acetophenone GC ∼1 g L−1 h−1 362
Vanadium chloroperoxidase 4-Pentenoic acid Bromolactone GC 1.4 g (∼80% yield) in 24 h 363
CPO-ILEMB/rGO/PEI/carbon cloth Phenol o-Chlorophenol and p-chlorophenol HPLC 364


6.1 Oxidation of alcohols, aldehydes, and related compounds

HRP can catalyze oxidation of organic compounds like phenols, alcohols, aldehydes, and chlorinated organics.60,365–372 HRP can be covalently immobilized on beads. The covalently bound HRP can be activated by H2O2 produced electrochemically to catalyze the oxidation of organic compounds. For example, HRP immobilized on beads via aminopropylation with 3-aminopropyltriethoxysilane and glutaraldehyde can oxidize phenol derivatives in an electrochemical reactor.348In situ electro-generation of H2O2 by reduction of dissolved oxygen was carried out using Pt coated on a Ti plate as the anode and stainless steel as the cathode. To generate the hydrogen peroxide, initially protons were generated by anodic oxidation of water (2H2O →O2 + 4H+ + 4e) and combined with dissolved oxygen to produce hydrogen peroxide at the cathode (O2 + 2H+ + 2e → H2O2). The aqueous phase of phenol was converted to p-benzoquinone, other organic acids, and carbon dioxide. The removal efficiency of the phenol reaction was 93%.

Combining peroxidases with electrochemistry can be applied to remove environmental hazards in biochemical processes. For example, bisphenol-A has biological toxicity to aquatic organisms and humans. HRP has high activity for oxidizing bisphenol-A in the presence of H2O2. Bioelectrocatalytic oxidation of bisphenol-A by HRP was done in a membrane-less electrochemical reactor.373 Silk fibroin (a natural fibrous protein), Fe3O4 nanoparticles and poly(amido amine) were dispersed in a phosphate buffer (pH = 7.4, 50 mM) for 8 h to assemble into magnetic composite nanoparticles (∼50 nm). Those magnetic nanoparticles were filtered and transferred to phosphate buffer containing HRP and 25% glutaraldehyde. After incubation at 4 °C for 8 h, the resulting magnetic particles decorated with HRP were isolated magnetically. In 0.1 M PBS buffer, H2O2 was produced by oxygen reduction on a carbon fiber cathode. The removal efficiency of bisphenol-A was 80% with an applied potential of 1.6 V, pH 5.0, and 25 mL min−1 oxygen flow rate at 25 °C.

Chloroperoxidase (CPO) is a fungal heme-thiolate enzyme that also has peroxidase activity. CPO is negatively charged, and it can assemble with polycations. For example, CPO and PDDA could assemble on a 3-mercaptopropanesulfonic acid-coated Au electrode. These PDDA-CPO-PDDA films were stable and retained 97% activity after storage for 48 h in pH 5 buffer.374 CPO films can electrochemically oxidize cinnamyl alcohol.356 CPO could be assembled on a DDAB/Nafion film on a GC electrode for bioelectrocatalytic oxidation of cinnamyl alcohol into cinnamyl aldehyde (Fig. 21A). GC modified with a Nafion film could hold a DDAB-CPO film after coating with chitosan. A well-defined, quasi-reversible redox pair for CPO at −0.2 V vs. SCE (phosphate buffer, pH 4.5) was obtained (Fig. 21B, curve a) with an electron transfer rate constant (ks) of 2.3 s−1. CPO reduced oxygen electrochemically to H2O2. Under saturated oxygen, a catalytic reduction peak was assigned to the production of H2O2 by CPO in the film as shown in Fig. 21B (curve b). H2O2 generated in situ by electrochemical reduction further drives the catalytic oxidation of cinnamyl alcohol, also catalyzed by CPO in the composite films. Cinnamyl aldehyde was formed at a yield of 52% after electrolysis for 2 h at −0.6 V vs. SCE. The product formation was confirmed by GC-MS.


image file: d3cs00461a-f21.tif
Fig. 21 (A) Oxidation of cinnamyl alcohol. (B) CVs of the CS-DDAB-CPO-Nafion/GCE (a) in nitrogen-saturated and (b) in oxygen-saturated phosphate buffer solution (50 mM, pH 4.5) at 100 mV s−1 scan rate. Reproduced from ChemCatChem, 2012, 4, 1850–1855 with permission.356 Copyright, 2012, John Wiley & Son.

Lignin peroxidase (LPO) activated by H2O2 can catalyze oxidative degradation of organic biopolymers, nonphenolic veratryl-type compounds, and a wide range of organic pollutants, such as phenol and non-phenolic compounds.47,375–380 Lee and Moon reported the oxidation of veratryl alcohol into veratraldehyde catalyzed in by LPO in solution with H2O2 generated by electrochemical reduction of oxygen.357 H2O2 was produced on a reticulated vitreous carbon cathode. The formation rate of veratraldehyde was 30 μM min−1 at −0.4 V vs. Ag/AgCl in a broad pH range of 2.5–5.5. The concentration of the veratraldehyde produced during the reaction was measured by UV-vis.

6.2 Oxidation of sulfur-containing compounds

Peroxidases can oxidize organo-sulfur compounds like thiols, sulfides and disulfides.381–384 The oxidation of aromatic sulfides has been reported by using LPO, SBP and CPO. The yield of sulfoxide decreased with electron donating power of substituents.385 Manganese peroxidase was reported to cleave the sulfide bond in di(2-methylpent-2-enyl) sulfide.386 The oxidation of omeprazole sulfide to (S)-omeprazole was catalyzed by SBP in CTAB/isooctane/n-butyl alcohol/water water-in-oil microemulsions.387 A bioelectrocatalytic system was developed with CPO for oxidation of thioanisole to (R)-methylphenylsulfoxide (Fig. 22).341 H2O2 generated by electroreduction of dissolved oxygen at the cathode could activate CPO to oxidize thioanisole. The enantioselectivity was >98% for R-methylphenylsulfoxide. The yield was 30 g L−1 d−1 at −0.5 V vs. Ag/AgCl in 300 mL sodium citrate buffer containing 30 vol% of t-butanol at pH 5. Similarly, the conversion of thioanisole into (R)-methylphenylsulfoxide was performed by using CPO.358 The electrolysis at a constant current produced H2O2 (eqn (11) and (12)). The product was (R)-methylphenylsulfoxide at formation rate 104 g L−1 d−1 as confirmed by HPLC and the product up to 1.2 g (enantiomeric excess >98.5%, purity >98%) was obtained.
image file: d3cs00461a-f22.tif
Fig. 22 A schematic representation of electroenzymatic sulfoxidation with chloroperoxidase.

The oxidation of methyl p-tolyl sulfide, N-MOC-L-methionine methyl ester and 1-methoxy-4-(methylthio)benzene into methyl p-tolyl sulfoxide, N-MOC-L-methionine methyl ester sulfoxide and methoxyphenylmethylsulfoxide was catalyzed by CPOs.359 Using a similar approach, H2O2 was generated electrochemically by reduction of oxygen at −0.5 V vs. Ag/AgCl. Under saturated oxygen, an irreversible reduction peak was found for reduction of oxygen to hydrogen peroxide between −0.5 V to −1.0 V vs. Ag/AgCl. The product conversion was 60% for N-MOC-L-methionine methyl ester, 76% for methyl p-tolyl sulfide, and 83% for 1-methoxy-4-(methylthio)benzene. The products of the reaction were identified by GC and NMR.

Cathode reaction:

 
O2 + 2H+ + 2e → H2O2(11)
Anode reaction:
 
H2O → 2e + 1/2O2 + 2H+(12)
Gas-diffusion electrodes have also been used for the peroxidase-catalyzed reactions. Gas-diffusion electrodes provide high concentrations of H2O2 in the vicinity of the enzyme.388 These gas-diffusion electrodes consisted of solid, liquid and gaseous interfaces and catalyst (e.g., carbon materials or carbon particles) for electrochemical reaction between liquid and gaseous phase, polytetrafluoroethylene (PTFE) and additives. PTFE provided a hydrophobic matrix to bind the catalyst and liquid attracting materials. Hydrogen peroxide produced at gas–liquid–solid interface diffuses into the liquid phase to the enzyme. For example, the gas-diffusion electrode was applied for enzymatic sulfoxidation, chlorination and oxidation by using CPO.360 The turnover numbers of enzymatic conversions of thioanisole to methyl phenyl sulfoxide, monochlorodimedone to dichlorodimedone, and indole to oxindole could reach 3120, 203[thin space (1/6-em)]100 and 39[thin space (1/6-em)]000, respectively. The reaction products were identified by GC-MS. The space time yields (STYs) for each reaction were observed 23, 9.5 and 8.3 g L−1 d−1, respectively (Fig. 23).


image file: d3cs00461a-f23.tif
Fig. 23 Biocatalytic reactions catalyzed by chloroperoxidase.

6.3 Hydroxylation

Selective hydroxylation of C–H bonds is typically slow with peroxidases.389 CPO was used for selective bromo-hydroxylation of alkenes,390 halo-hydroxylation of monoterpenes,391 and hydroxylation of C–H bonds in ethylbenzene (Fig. 24).361 Asymmetric hydroxylation of ethylbenzene to R-1-phenylethanol was carried out in the presence of in situ H2O2 generation by oxidation of monosaccharides catalyzed by gold nanoclusters (AuNCs). AuNCs were incorporated into mesoporous silica with hydroxyl groups (HMS-OH), denoted Au@HMS. CPO was loaded on the surface of Au@HMS (CPO-Au@HMS) and placed into the reaction solution with acetate buffer (pH 5.5), ethylbenzene and galactose for 12 h at 30 °C. The oxidation of galactose by AuNCs produced a high concentration of H2O2 (1.2 mM) that further activates CPO. The yield of R-1-phenylethanol was ∼1.6 mM in 12 h when 120 mM of galactose was used in this consecutive reaction.
image file: d3cs00461a-f24.tif
Fig. 24 Hydroxylation of C–H bond using chloroperoxidase to convert ethyl benzene into 1-phenethyl alcohol. Reproduced from ChemCatChem, 2022, 14(4), e202101732 with permission.361 Copyright, 2022, John Wiley & Son.

In recent years, unspecific peroxygenases (UPOs) containing heme-thiolate have attracted much attention to catalyze monooxygenation of various organic compounds with H2O2. UPOs can be isolated from the natural fungus, e.g., the edible mushroom Agrocybe aegerita.392,393 Similar to other peroxidases, UPOs do not catalyze the reductive activation of oxygen, but it directly uses H2O2 to generate the catalytic active oxyferryl.392 An excess of H2O2, however, can decrease peroxygenase activity; and in some cases, it can drive over oxidation which leads to low turnover and byproduct formation. Electroenzymatic hydroxylation of ethylbenzene to 1-phenethyl alcohol with UPOs produced by fungus Agrocybe aegerita (AaeUPO) has been demonstrated by Horst et al.362 Oxygen was used from ambient air and protons from the buffer for electrochemical in situ generation of H2O2 on a carbon black gas diffusion electrode. The current efficiency was 65–78% in 30 mL of reaction mixtures containing AaeUPO (50 nM) with phosphate buffer (100 mM, pH, 7.0) containing 3 vol% of acetone and 500 μL of ethylbenzene. The product was quantified on a GC. The highest turnover number was 400[thin space (1/6-em)]000 at −10 mA cm−2 after 4 h electrolysis and the space-time-yield was up to ∼1 g L−1 h−1. However, the AaeUPO was non-specific to produce 1-phenethyl alcohol and 10–20% of acetophenone as an overoxidized byproduct was seen under constant-current electrolysis.

6.4 Halogenation

CPO and vanadium peroxidase are active for enzymatic halogenation.360,363,394 CPO can catalyze the halogenation by using free halide ions and H2O2. Bioelectrochemical synthesis of chlorinated barbituric acid was studied by using CPO.395 H2O2 was produced in an electrolytic cell at −0.5 V vs. SCE and activates CPO in a membrane chamber to produce 5-chlorobarbituric acid. The conversion of barbituric acid into 5-chlorobarbituric acid was >96% after 24 h. The enzymatic conversion of 4-pentenoic acid to bromolactone was catalyzed by vanadium chloroperoxidase with in situ generation of H2O2.363 An oxidized carbon nanotube (o-CNT) gas-diffusion electrode was used as a cathode where H2O2 was produced from oxygen reduction. Vanadium chloroperoxidase was used in solution for conversion of 4-pentenoic acid to bromolactone product identified by GC. The isolated yield of bromolactone was 1.4 g (∼80% yield) in 24 h. Chlorination of phenol was demonstrated with ionic liquid modified CPO (CPO-ILEMB) assembled on a reduced graphene oxide electrode (rGO)/PEI (rGO/PEI).364 The enzymatic reaction was performed in the range −0.3 V to −0.7 V vs. SCE with chloride ions (Cl) as the source for chlorination. The composite of rGO/PEI and CPO-ILEMB catalyzed ortho-selective phenol chlorination (Fig. 25). At −0.3 V to −0.7 V, the yield of ortho-chlorophenol increased with potential. Two products, o-chlorophenol and p-chlorophenol, were found only at −0.6 V vs. SCE. The protocol was selective at −0.3 V and −0.4 V for synthesis of o-chlorophenol. Catalytic efficiency was improved 5.5-fold as compared to manually added H2O2 and 2.3-fold compared to the use of free CPO. The products were confirmed by HPLC.
image file: d3cs00461a-f25.tif
Fig. 25 (a) Scheme for fabrication of CPO-ILEMB/rGO/PEI/carbon cloth electrode and (b) electroenzymatic protocol on carbon cloth to synthesis of ortho-chlorophenol. Reproduced from ACS Sustainable Chem. Eng., 10, 12497–12503 with permission.364 Copyright, 2022, ACS Publications.

7. Conclusions and prospective

Selective oxidations are of critical importance in organic and pharmaceutical syntheses. Effective bioelectrocatalytic approaches using heme-containing enzymes, especially Cyt P450s and peroxidases, have been demonstrated to serve this purpose as described above, and have attracted tremendous interest. While bioelectrocatalytic reactions are important as a synthetic tool, a major goal is to refine the methodologies to generate products with high yields and high selectivity. The practical utility of bioelectrocatalysis is heavily impacted by the stability of enzymes, and selected immobilization methods can achieve high stability and large turnover numbers. For example, crosslinking enzymes into thin polyion films enables high enzyme stability and reaction chemistry at temperatures well above the solution decomposition temperature of the enzyme and provide much higher reaction rates at much higher temperatures than those where unmodified enzymes are stable in solution.

Key challenges with enzymes include thermodynamics, kinetics, and non-aqueous solvent stability (e.g., in organic solvents). The first two aspects have been addressed well by crosslinking and other film strategies, while non-aqueous solvents can be used in microemulsions that do not denature enzymes in films. In addition, genetic molecular engineering by site-directed mutagenesis of enzymes offers a pathway to high thermodynamic and kinetic stability as well as improved reactivity of enzymes, some resistance to organic solvents, accessibility to new or improved reactions, and alternative product selectivity.396,397 These genetically engineered enzymes can play a significant role for highly specific and selective organic synthesis. For example, Arnold et al. reported a number of engineered heme-containing enzymes, e.g., P450BM3 as high efficient and specific for C–H activation.398 In a recent study, engineered P450 nitrene transferase was demonstrated for conversion of aliphatic C–H bonds to chiral amines and cyclohexane C–H bonds to acetamide.399 These engineered enzymes have not been used much in bioelectrocatalysis, but could offer effective pathways to improve the specificity and access to organic molecules that natural enzymes do not catalyze. However, the cost of genetically engineered enzyme production is an important issue compared to natural enzymes that are often easily obtained from natural sources.

As mentioned above, immobilizing enzymes in stable films, e.g., crosslinked LbL films using polyions with amine groups, offers excellent promise for high stability and retention of activity under synthetic conditions at relatively high temperatures. This has been demonstrated for peroxidases up to 90 °C in aqueous buffer and in microemulsions. More attention should be paid to electrocatalysis in microemulsions, which provides solvation of water-insoluble hydrophobic reactants, as well as the sufficient water to hydrate enzymes in films and allow them to function in a close-to-natural environment at a relatively low cost.

As discussed, for Cyt P450s in microsomes or supersomes, electron transfer proceeds from electrodes to Cyt P450 reductase (CPR), then to the Cyt P450 FeIIIheme, and subsequently formed species utilize molecular oxygen directly as in the natural pathway (Scheme 2). For peroxidases, however, an external source of peroxide is required to activate the enzyme. In situ, on-demand, controlled generation of H2O2 is attractive for this purpose to enhance atom economy, yields, and enzyme stability. Thus, more feasible and effective total systems are required to provide hydrogen peroxide using electrochemical or alternative approaches. For example, complete flow reactor systems that catalytically produce and biocatalyze synthetic reactions have so far seen little attention. Photocatalytic production of H2O2 from O2 is also a viable option and could be combined with an electrochemical cell to activate enzymes. In our opinion, FeIIIheme enzymes are the most stable, versatile, and best utilized for biosynthetic applications in thin films on an electrode or on nanoparticles with electrochemical generation of activators such as H2O2. In view of scaleup, new designs of effective reaction cells and protocols which can produce product at larger scale or in a high-throughput fashion are very important future goals.

Finally, bioelectrocatalytic synthesis is limited somewhat by the numbers and types of enzymes available. Biological reactions are often catalyzed by a group of enzymes in sequence, and this approach may also have a role in chemical syntheses. A good example of such chemoenzymatic pathways was reported by Cai et al.400 who demonstrated the enzymatic conversion of CO2 into starch using 11 sequential reactions. The surface of electrodes in bioelectrochemical synthesis indeed can support the formation of multienzyme complexes with spatial proximity to catalyze cascade reactions, but there are few reports of bioelectrocatalytic syntheses involving consecutive reactions. Alternatively, using synthetic catalysts to replace some recalcitrant enzyme steps in multienzyme syntheses could also be a viable alternative for scaleup. In a recent study, an electro-biosystem was reported using nanostructured copper to convert CO2 to acetate and genetically engineered Saccharomyces cerevisiae to then produce glucose from the acetate.401 Integrating synthetic catalysts with multienzyme syntheses may be able to improve overall reaction efficiency and could broaden synthetic capacity and the applications of bioelectrocatalytic synthesis.

Conflicts of interest

The authors declare no conflict of interest.

Acknowledgements

The authors are grateful for financial support from a National Science Foundation grant (CBET-2035669) and partial support from the University of Connecticut Research Excellence Program.

References

  1. J. J. Sangster, J. R. Marshall, N. J. Turner and J. Mangas-Sanchez, ChemBioChem, 2022, 23, e202100464 CrossRef CAS PubMed.
  2. R. Siedentop, C. Claaßen, D. Rother, S. Lütz and K. Rosenthal, Catalysts, 2021, 11, 1183 CrossRef CAS.
  3. M. Ghaffari-Moghaddam, H. Eslahi, Y. A. Aydin and D. Saloglu, J. Biol. Methods, 2015, 2, e25 CrossRef.
  4. E. L. Bell, W. Finnigan, S. P. France, A. P. Green, M. A. Hayes, L. J. Hepworth, S. L. Lovelock, H. Niikura, S. Osuna, E. Romero, K. S. Ryan, N. J. Turner and S. L. Flitsch, Nat. Rev. Methods Primer, 2021, 1, 46 CrossRef CAS.
  5. K. M. Koeller and C.-H. Wong, Nature, 2001, 409, 232–240 CrossRef CAS PubMed.
  6. R. P. Sheridan, K. R. Korzekwa, R. A. Torres and M. J. Walker, J. Med. Chem., 2007, 50, 3173–3184 CrossRef CAS PubMed.
  7. M. Zhao, J. Ma, M. Li, Y. Zhang, B. Jiang, X. Zhao, C. Huai, L. Shen, N. Zhang, L. He and S. Qin, Int. J. Mol. Sci., 2021, 22, 12808 CrossRef CAS PubMed.
  8. F. P. Guengerich, Chem. Res. Toxicol., 2001, 14, 611–650 Search PubMed.
  9. S. M. Mousavi, S. A. Hashemi, S. M. Iman Moezzi, N. Ravan, A. Gholami, C. W. Lai, W.-H. Chiang, N. Omidifar, K. Yousefi and G. Behbudi, Biochem. Res. Int., 2021, 2021, 1–12 Search PubMed.
  10. S. Rathore, A. Varshney, S. Mohan and P. Dahiya, Biotechnol. Genet. Eng. Rev., 2022, 38, 1–32 Search PubMed.
  11. J. Gan, M. Bilal, X. Li, S. Z. Hussain Shah, B. A. Mohamed, T. Hadibarata and H. Cheng, Chemosphere, 2022, 307, 136035 CrossRef CAS PubMed.
  12. Cytochrome P450: Structure, Mechanism, and Biochemistry, ed. P. R. O. de Montellano, Springer US, Boston, MA, 1995 Search PubMed.
  13. Cytochrome P450, ed. P. R. Ortiz de Montellano, Springer US, Boston, MA, 2005 Search PubMed.
  14. P. Anzenbacher and E. Anzenbacherová, Cell. Mol. Life Sci., 2001, 58, 737–747 CrossRef CAS PubMed.
  15. L. L. von Moltke, D. J. Greenblatt, J. Schmider, J. S. Harmatz and R. I. Shader, Clin. Pharmacokinet., 1995, 29, 33–44 CrossRef CAS PubMed.
  16. A. M. McDonnell and C. H. Dang, J. Adv. Pract. Oncol., 2013, 4, 263–268 Search PubMed.
  17. T. Lynch and A. Price, Cytochrome P, 2007, 76, 6 Search PubMed.
  18. S. Uehara, Y. Uno and H. Yamazaki, Biochem. Pharmacol., 2020, 171, 113721 CrossRef CAS PubMed.
  19. U. M. Zanger and M. Schwab, Pharmacol. Ther., 2013, 138, 103–141 CrossRef CAS PubMed.
  20. S. Li, L. Du and R. Bernhardt, Trends Microbiol., 2020, 28, 445–454 CrossRef CAS PubMed.
  21. M. J. Cryle, J. E. Stok and J. J. De Voss, Aust. J. Chem., 2003, 56, 749 CrossRef CAS.
  22. S. Cholerton, A. K. Daly and J. R. Idle, Trends Pharmacol. Sci., 1992, 13, 434–439 CrossRef CAS PubMed.
  23. M. Spatzenegger and W. Jaeger, Drug Metab. Rev., 1995, 27, 397–417 CrossRef CAS PubMed.
  24. E. Morgan, Clin. Pharmacol. Ther., 2009, 85, 434–438 CrossRef CAS PubMed.
  25. E. G. Hvastkovs, J. B. Schenkman and J. F. Rusling, Annu. Rev. Anal. Chem., 2012, 5, 79–105 CrossRef CAS PubMed.
  26. E. G. Hvastkovs and J. F. Rusling, Anal. Chem., 2016, 88, 4584–4599 CrossRef CAS PubMed.
  27. J. F. Rusling and E. G. Hvastkovs, Drug Metabolism Handbook, Wiley, 2022, vol. 2 Search PubMed.
  28. S. B. Lamb, D. C. Lamb, S. L. Kelly and D. C. Stuckey, FEBS Lett., 1998, 431, 343–346 CrossRef CAS PubMed.
  29. T. Sakaki, K. Yamamoto and S. Ikushiro, Biotechnol. Appl. Biochem., 2013, 60, 65–70 CrossRef CAS PubMed.
  30. J. B. Y. H. Behrendorff, Front. Microbiol., 2021, 12, 649273 CrossRef PubMed.
  31. M. A. Alonso-Lomillo, J. Gonzalo-Ruiz, O. Domínguez-Renedo, F. J. Muñoz and M. J. Arcos-Martínez, Biosens. Bioelectron., 2008, 23, 1733–1737 CrossRef CAS PubMed.
  32. L. Asturias-Arribas, M. A. Alonso-Lomillo, O. Domínguez-Renedo and M. J. Arcos-Martínez, Anal. Chim. Acta, 2011, 685, 15–20 CrossRef CAS PubMed.
  33. P. Sun and Y. Wu, Sens. Actuators B Chem., 2013, 178, 113–118 CrossRef CAS.
  34. J. Ducharme and K. Auclair, Biochim. Biophys. Acta, Proteins Proteomics, 2018, 1866, 32–51 CrossRef CAS PubMed.
  35. R. W. Estabrook, K. M. Faulkner, M. S. Shet and C. W. Fisher, Methods in Enzymology, Elsevier, 1996, vol. 272, pp. 44–51 Search PubMed.
  36. L. Zhang, Z. Xie, Z. Liu, S. Zhou, L. Ma, W. Liu, J.-W. Huang, T.-P. Ko, X. Li, Y. Hu, J. Min, X. Yu, R.-T. Guo and C.-C. Chen, Nat. Commun., 2020, 11, 2676 CrossRef CAS PubMed.
  37. K. J. McLean, D. Luciakova, J. Belcher, K. L. Tee and A. W. Munro, in Monooxygenase, Peroxidase and Peroxygenase Properties and Mechanisms of Cytochrome P450, ed. E. G. Hrycay and S. M. Bandiera, Springer International Publishing, Cham, 2015, vol. 851, pp. 299–317 Search PubMed.
  38. B. Meunier, S. P. de Visser and S. Shaik, Chem. Rev., 2004, 104, 3947–3980 CrossRef CAS PubMed.
  39. J. T. Groves, J. Inorg. Biochem., 2006, 100, 434–447 CrossRef CAS PubMed.
  40. F. P. Guengerich, ACS Catal., 2018, 8, 10964–10976 CrossRef CAS PubMed.
  41. J. F. Rusling, E. G. Hvastkovs, D. O. Hull and J. B. Schenkman, Chem. Commun., 2008, 141–154 RSC.
  42. A. Claesson and O. Spjuth, Mini-Rev. Med. Chem., 2013, 13, 720–729 CrossRef CAS PubMed.
  43. T. L. de Albuquerque, J. S. Silva, A. C. de Macedo, L. R. B. Gonçalves and M. V. P. Rocha, Reference Module in Chemistry, Molecular Sciences and Chemical Engineering, Elsevier, 2019 Search PubMed.
  44. S. Colonna, N. Gaggero, C. Richelmi and P. Pasta, Trends Biotechnol., 1999, 17, 163–168 CrossRef CAS PubMed.
  45. K. N. Ingenbosch, S. Quint, M. Dyllick-Brenzinger, D. S. Wunschik, J. Kiebist, P. Süss, U. Liebelt, R. Zuhse, U. Menyes, K. Scheibner, C. Mayer, K. Opwis, J. S. Gutmann and K. Hoffmann-Jacobsen, ChemBioChem, 2021, 22, 398–407 CrossRef CAS PubMed.
  46. S. Chen and P. Schopfer, Eur. J. Biochem., 1999, 260, 726–735 CrossRef CAS PubMed.
  47. A. O. Falade, U. U. Nwodo, B. C. Iweriebor, E. Green, L. V. Mabinya and A. I. Okoh, MicrobiologyOpen, 2017, 6, e00394 CrossRef PubMed.
  48. M. Tien and D. Ma, J. Biol. Chem., 1997, 272, 8912–8917 CrossRef CAS PubMed.
  49. M. Tanaka, K. Matsuura, S. Yoshioka, S. Takahashi, K. Ishimori, H. Hori and I. Morishima, Biophys. J., 2003, 84, 1998–2004 CrossRef CAS PubMed.
  50. E. G. Hrycay and S. M. Bandiera, in Monooxygenase, Peroxidase and Peroxygenase Properties and Mechanisms of Cytochrome P450, ed. E. G. Hrycay and S. M. Bandiera, Springer International Publishing, Cham, 2015, vol. 851, pp. 1–61 Search PubMed.
  51. D. C. Goodwin, S. D. Aust and T. A. Grover, Biochemistry, 1995, 34, 5060–5065 CrossRef CAS PubMed.
  52. V. Christian, R. Shrivastava, D. Shukla, H. Modi and B. R. M. Vyas, Enzyme Microb. Technol., 2005, 36, 327–332 CrossRef CAS.
  53. R. S. Koduri and M. Tien, J. Biol. Chem., 1995, 270, 22254–22258 CrossRef CAS PubMed.
  54. E. Vignali, F. Tonin, L. Pollegioni and E. Rosini, Appl. Microbiol. Biotechnol., 2018, 102, 10579–10588 CrossRef CAS PubMed.
  55. N. Khatoon, A. Jamal and M. I. Ali, Environ. Technol., 2019, 40, 1366–1375 CrossRef CAS PubMed.
  56. J. Guo, X. Liu, X. Zhang, J. Wu, C. Chai, D. Ma, Q. Chen, D. Xiang and W. Ge, Int. J. Biol. Macromol., 2019, 138, 433–440 CrossRef CAS PubMed.
  57. N. C. Veitch, Phytochemistry, 2004, 65, 249–259 CrossRef CAS PubMed.
  58. M. Aldhahri, Y. Q. Almulaiky, R. M. El-Shishtawy, W. M. Al-Shawafi, N. Salah, A. Alshahrie and H. A. H. Alzahrani, Catal. Lett., 2021, 151, 232–246 CrossRef CAS.
  59. J. P. McEldoon and J. S. Dordick, J. Biol. Chem., 1991, 266, 14288–14293 CrossRef CAS PubMed.
  60. M. Jonović, B. Jugović, M. Žuža, V. Đorđević, N. Milašinović, B. Bugarski and Z. Knežević-Jugović, Polymers, 2022, 14, 2614 CrossRef PubMed.
  61. M. Pešić, C. López, G. Álvaro and J. López-Santín, J. Mol. Catal. B: Enzym., 2012, 84, 144–151 CrossRef.
  62. A. But, A. van Noord, F. Poletto, J. P. M. Sanders, M. C. R. Franssen and E. L. Scott, Mol. Catal., 2017, 443, 92–100 CrossRef CAS.
  63. E. Kiljunen and L. T. Kanerva, J. Mol. Catal. B: Enzym., 2000, 9, 163–172 CrossRef CAS.
  64. G. Masdeu, M. Pérez-Trujillo, J. López-Santín and G. Álvaro, Process Biochem., 2016, 51, 1204–1211 CrossRef CAS.
  65. A. D’Annibale, C. Crestini, E. D. Mattia and G. G. Sermanni, J. Biotechnol., 1996, 48, 231–239 CrossRef.
  66. P. Chowdhary, G. Shukla, G. Raj, L. F. R. Ferreira and R. N. Bharagava, SN Appl. Sci., 2019, 1, 45 CrossRef.
  67. J. M. Naapuri, P. K. Wagner, F. Hollmann and J. Deska, ChemistryOpen, 2022, 11, e2021002 CrossRef PubMed.
  68. Md Belal, S. Sarkar, R. Subramanian and A. T. Khan, Org. Biomol. Chem., 2022, 20, 2562–2579 RSC.
  69. Q. Luo, J. Li, W. Wang, Y. Li, Y. Li, X. Huo, J. Li and N. Wang, ACS Appl. Mater. Interfaces, 2022, 14, 14218–14225 CrossRef CAS PubMed.
  70. P. Dhankhar, V. Dalal, V. Singh, A. K. Sharma and P. Kumar, Int. J. Biol. Macromol., 2021, 193, 601–608 CrossRef CAS PubMed.
  71. J. He, Y. Zhang, Q. Yuan and H. Liang, Ind. Eng. Chem. Res., 2019, 58, 3555–3560 CrossRef CAS.
  72. F. Gao, L. Wang, Y. Liu, S. Wang, Y. Jiang, M. Hu, S. Li and Q. Zhai, Biochem. Eng. J., 2015, 93, 243–249 CrossRef CAS.
  73. B. Valderrama, M. Ayala and R. Vazquez-Duhalt, Chem. Biol., 2002, 9, 555–565 CrossRef CAS PubMed.
  74. H. L. Wapshott-Stehli and A. M. Grunden, Enzyme Microb. Technol., 2021, 145, 109744 CrossRef CAS PubMed.
  75. S. Ortega Ugalde, C. P. de Koning, K. Wallraven, B. Bruyneel, N. P. E. Vermeulen, T. N. Grossmann, W. Bitter, J. N. M. Commandeur and J. C. Vos, Appl. Microbiol. Biotechnol., 2018, 102, 9231–9242 CrossRef CAS PubMed.
  76. S. B. Mellor, A. Z. Nielsen, M. Burow, M. S. Motawia, D. Jakubauskas, B. L. Møller and P. E. Jensen, ACS Chem. Biol., 2016, 11, 1862–1869 CrossRef CAS PubMed.
  77. L. Waskell and J.-J. P. Kim, in Cytochrome P450, ed. P. R. Ortiz de Montellano, Springer International Publishing, Cham, 2015, pp. 33–68 Search PubMed.
  78. M. J. I. Paine, N. S. Scrutton, A. W. Munro, A. Gutierrez, G. C. K. Roberts and C. R. Wolf, in Cytochrome P450, ed. P. R. Ortiz de Montellano, Springer US, Boston, MA, 2005, pp. 115–148 Search PubMed.
  79. Z. E. Chiliza, J. Martínez-Oyanedel and K. Syed, Biophys. Rev., 2020, 12, 1217–1222 CrossRef CAS PubMed.
  80. S. A. Child, J. M. Bradley, T. L. Pukala, D. A. Svistunenko, N. E. Le Brun and S. G. Bell, Chem. Sci., 2018, 9, 7948–7957 RSC.
  81. M. Sasaki, A. Akahira, K. Oshiman, T. Tsuchido and Y. Matsumura, Appl. Environ. Microbiol., 2005, 71, 8024–8030 CrossRef CAS PubMed.
  82. D. C. Lamb, Y. Kim, L. V. Yermalitskaya, V. N. Yermalitsky, G. I. Lepesheva, S. L. Kelly, M. R. Waterman and L. M. Podust, Structure, 2006, 14, 51–61 CrossRef CAS PubMed.
  83. A. W. Munro, H. M. Girvan and K. J. McLean, Biochim. Biophys. Acta, Gen. Subj., 2007, 1770, 345–359 CrossRef CAS PubMed.
  84. S. Bellosta, R. Paoletti and A. Corsini, Circulation, 2004, 109, III-50–III-57 CrossRef PubMed.
  85. O. V. Maksymchuk and V. I. Kashuba, Pharmacol. Rep., 2020, 72, 1161–1172 CrossRef CAS PubMed.
  86. H. Hashimoto, K. Toide, R. Kitamura, M. Fujita, S. Tagawa, S. Itoh and T. Kamataki, Eur. J. Biochem., 1993, 218, 585–595 CrossRef CAS PubMed.
  87. F. P. Guengerich, in Cytochrome P450, ed. P. R. Ortiz de Montellano, Springer International Publishing, Cham, 2015, pp. 523–785 Search PubMed.
  88. F. P. Guengerich, in Cytochrome P450, ed. P. R. Ortiz de Montellano, Springer US, Boston, MA, 2005, pp. 377–530 Search PubMed.
  89. M. R. McGill and H. Jaeschke, Pharm. Res., 2013, 30, 2174–2187 CrossRef CAS PubMed.
  90. H.-J. Xie, Ü. Yasar, S. Lundgren, L. Griskevicius, Y. Terelius, M. Hassan and A. Rane, Pharmacogenomics J., 2003, 3, 53–61 CrossRef CAS PubMed.
  91. P. Roy, L. J. Yu, C. L. Crespi and D. J. Waxman, Drug Metab. Dispos., 1999, 27, 655–666 CAS.
  92. D. D. Bikle, Chem. Biol., 2014, 21, 319–329 CrossRef CAS PubMed.
  93. G. Jones, D. E. Prosser and M. Kaufmann, J. Lipid Res., 2014, 55, 13–31 CrossRef CAS PubMed.
  94. I. Aiba, T. Yamasaki, T. Shinki, S. Izumi, K. Yamamoto, S. Yamada, H. Terato, H. Ide and Y. Ohyama, Steroids, 2006, 71, 849–856 CrossRef CAS PubMed.
  95. O. Pelkonen, M. Turpeinen, J. Hakkola, P. Honkakoski, J. Hukkanen and H. Raunio, Arch. Toxicol., 2008, 82, 667–715 CrossRef CAS PubMed.
  96. X. Qin and X. Wang, Acta Pharm. Sin. B, 2019, 9, 1087–1098 CrossRef PubMed.
  97. S. Hiraga, K. Sasaki, H. Ito, Y. Ohashi and H. Matsui, Plant Cell Physiol., 2001, 42, 462–468 CrossRef CAS PubMed.
  98. K. G. Welinder, Curr. Opin. Struct. Biol., 1992, 2, 388–393 CrossRef CAS.
  99. V. S. Ferreira-Leitão, J. G. da Silva and E. P. S. Bon, Appl. Catal., 2003, 42, 213–221 CrossRef.
  100. M. Zámocký, P. G. Furtmüller and C. Obinger, Arch. Biochem. Biophys., 2010, 500, 45–57 CrossRef PubMed.
  101. F. Passardi, G. Theiler, M. Zamocky, C. Cosio, N. Rouhier, F. Teixera, M. Margis-Pinheiro, V. Ioannidis, C. Penel, L. Falquet and C. Dunand, Phytochemistry, 2007, 68, 1605–1611 CrossRef CAS PubMed.
  102. N. Bistolas, U. Wollenberger, C. Jung and F. W. Scheller, Biosens. Bioelectron., 2005, 20, 2408–2423 CrossRef CAS PubMed.
  103. A. Yarman, U. Wollenberger and F. W. Scheller, Electrochim. Acta, 2013, 110, 63–72 CrossRef CAS.
  104. V. A. Dixit, L. A. Lal and S. R. Agrawal, WIREs Comput. Mol. Sci., 2017, 7, e1323 Search PubMed.
  105. O. A. Fatunde and S.-A. Brown, Int. J. Mol. Sci., 2020, 21, 604 CrossRef CAS PubMed.
  106. D. Valikhani, J. M. Bolivar and J. N. Pelletier, ACS Catal., 2021, 11, 9418–9434 CrossRef CAS.
  107. J. B. Y. H. Behrendorff and E. M. J. Gillam, Chem. Res. Toxicol., 2017, 30, 453–468 Search PubMed.
  108. V. V. Shumyantseva, A. V. Kuzikov, R. A. Masamrekh, T. V. Bulko and A. I. Archakov, Biosens. Bioelectron., 2018, 121, 192–204 CrossRef CAS PubMed.
  109. B. D. Fleming, D. L. Johnson, A. M. Bond and L. L. Martin, Expert Opin. Drug Metab. Toxicol., 2006, 2, 581–589 CrossRef CAS PubMed.
  110. L. Asturias-Arribas, M. A. Alonso-Lomillo, O. Domínguez-Renedo and M. J. Arcos-Martínez, Talanta, 2013, 105, 131–134 CrossRef CAS PubMed.
  111. S. Kumar, Expert Opin. Drug Metab. Toxicol., 2010, 6, 115–131 CrossRef CAS PubMed.
  112. E. Schneider and D. S. Clark, Biosens. Bioelectron., 2013, 39, 1–13 CrossRef CAS PubMed.
  113. J. Genovino, D. Sames, L. G. Hamann and B. B. Touré, Angew. Chem. Int. Ed., 2016, 55, 14218–14238 CrossRef CAS PubMed.
  114. V. B. Urlacher and M. Girhard, Trends Biotechnol., 2019, 37, 882–897 CrossRef CAS PubMed.
  115. P. R. Ortiz de Montellano, in Cytochrome P450, ed. P. R. Ortiz de Montellano, Springer International Publishing, Cham, 2015, pp. 111–176 Search PubMed.
  116. S. Krishnan, J. B. Schenkman and J. F. Rusling, J. Phys. Chem. B, 2011, 115, 8371–8380 CrossRef CAS PubMed.
  117. G.-D. Roiban and M. T. Reetz, Chem. Commun., 2015, 51, 2208–2224 RSC.
  118. J. F. Rusling and Z. Zhang, Handbook of Surfaces and Interfaces of Materials, Elsevier, 2001, pp. 33–71 Search PubMed.
  119. U. Bussy, R. Boisseau, C. Thobie-Gautier and M. Boujtita, TrAC, Trends Anal. Chem., 2015, 70, 67–73 CrossRef CAS.
  120. U. Bussy, Y.-W. Chung-Davidson, K. Li and W. Li, Environ. Sci. Technol., 2015, 49, 4450–4457 CrossRef CAS PubMed.
  121. L. Portychová and K. A. Schug, Anal. Chim. Acta, 2017, 993, 1–21 CrossRef PubMed.
  122. J. Lu, D. Cui, H. Li, Y. Zhang and S. Liu, Electrochim. Acta, 2015, 165, 36–44 CrossRef CAS.
  123. F. W. Scheller, U. Wollenberger, C. Lei, W. Jin, B. Ge, C. Lehmann, F. Lisdat and V. Fridman, Rev. Mol. Biotechnol., 2002, 82, 411–424 CrossRef CAS PubMed.
  124. L. H. Mak, S. J. Sadeghi, A. Fantuzzi and G. Gilardi, Anal. Chem., 2010, 82, 5357–5362 CrossRef CAS PubMed.
  125. S. J. Sadeghi, S. Ferrero, G. Di Nardo and G. Gilardi, Bioelectrochemistry, 2012, 86, 87–91 CrossRef CAS PubMed.
  126. V. V. Shumyantseva, A. A. Makhova, E. V. Shikh, T. V. Bulko, A. V. Kuzikov, R. A. Masamrekh, T. Shkel, S. Usanov, A. Gilep and A. I. Archakov, Bionanoscience, 2019, 9, 79–86 CrossRef.
  127. M. Huang, X. Xu, H. Yang and S. Liu, RSC Adv., 2012, 2, 12844 RSC.
  128. S. J. Sadeghi, A. Fantuzzi and G. Gilardi, Biochim. Biophys. Acta, Proteins Proteomics, 2011, 1814, 237–248 CrossRef CAS PubMed.
  129. V. V. Shumyantseva, Y. D. Ivanov, N. Bistolas, F. W. Scheller, A. I. Archakov and U. Wollenberger, Anal. Chem., 2004, 76, 6046–6052 CrossRef CAS PubMed.
  130. S. Krishnan, A. Abeykoon, J. B. Schenkman and J. F. Rusling, J. Am. Chem. Soc., 2009, 131, 16215–16224 CrossRef CAS PubMed.
  131. S. Krishnan, D. Wasalathanthri, L. Zhao, J. B. Schenkman and J. F. Rusling, J. Am. Chem. Soc., 2011, 133, 1459–1465 CrossRef CAS PubMed.
  132. R. Nerimetla, G. Premaratne, H. Liu and S. Krishnan, Electrochim. Acta, 2018, 280, 101–107 CrossRef CAS.
  133. C. Carucci, A. Salis and E. Magner, Curr. Opin. Electrochem., 2017, 5, 158–164 CrossRef CAS.
  134. K. M. Faulkner, M. S. Shet, C. W. Fisher and R. W. Estabrook, Proc. Natl. Acad. Sci. U. S. A., 1995, 92, 7705–7709 CrossRef CAS PubMed.
  135. A. K. Udit, F. H. Arnold and H. B. Gray, J. Inorg. Biochem., 2004, 98, 1547–1550 CrossRef CAS PubMed.
  136. K. Drauz, H. Gröger and O. May, Enzyme Catalysis in Organic Synthesis, John Wiley & Sons, 2012, vol. 1 Search PubMed.
  137. S. M. Mense and L. Zhang, Cell Res., 2006, 16, 681–692 CrossRef CAS PubMed.
  138. F. Scott Mathews, Prog. Biophys. Mol. Biol., 1985, 45, 1–56 CrossRef PubMed.
  139. P. Ponka, Am. J. Med. Sci., 1999, 318, 241–256 CrossRef CAS PubMed.
  140. J. D. Dimitrov and T. L. Vassilev, Nat. Biotechnol., 2009, 27, 892 CrossRef CAS PubMed.
  141. Y. Xing, S. Gao, X. Zhang and J. Zang, Foods, 2022, 11, 3594 CrossRef CAS PubMed.
  142. W. Hummel and H. Gröger, J. Biotechnol., 2014, 191, 22–31 CrossRef CAS PubMed.
  143. M.-R. Kula and C. Wandrey, Methods in Enzymology, Elsevier, 1987, vol. 136, pp. 9–21 Search PubMed.
  144. Y. Wu, H. Li, X.-M. Zhang, J.-S. Gong, H. Li, Z.-M. Rao, J.-S. Shi and Z.-H. Xu, Steroids, 2015, 101, 15–20 CrossRef CAS PubMed.
  145. F. Hollmann and A. Schmid, Biocatal. Biotransform., 2004, 22, 63–88 CrossRef CAS.
  146. D. Valikhani, J. M. Bolivar, A. Dennig and B. Nidetzky, Biotechnol. Bioeng., 2018, 115, 2416–2425 CrossRef CAS PubMed.
  147. A. M. Chánique and L. P. Parra, Front. Microbiol., 2018, 9, 194 CrossRef PubMed.
  148. J. P. Gray, V. Mishin, D. E. Heck, D. L. Laskin and J. D. Laskin, Toxicol. Appl. Pharmacol., 2010, 247, 76–82 CrossRef CAS PubMed.
  149. R. Zangar, Toxicol. Appl. Pharmacol., 2004, 199, 316–331 CrossRef CAS PubMed.
  150. I. Hanukoglu, Drug Metab. Rev., 2006, 38, 171–196 CrossRef CAS PubMed.
  151. S.-Y. Park, K. Yamane, S. Adachi, Y. Shiro, K. E. Weiss, S. A. Maves and S. G. Sligar, J. Inorg. Biochem., 2002, 91, 491–501 CrossRef CAS PubMed.
  152. K. Nguyen, N. Nguyen, M. Milhim, V. Nguyen, T. Lai, H. Nguyen, T. Le, T. Phan and R. Bernhardt, FEBS Open Bio, 2021, 11, 124–132 CrossRef CAS PubMed.
  153. H. Matsumura, S. Wiwatchaiwong, N. Nakamura, M. Yohda and H. Ohno, Electrochem. Commun., 2006, 8, 1245–1249 CrossRef CAS.
  154. E. Blair, J. Greaves and P. J. Farmer, J. Am. Chem. Soc., 2004, 126, 8632–8633 CrossRef CAS PubMed.
  155. H. Matsumura, N. Nakamura, M. Yohda and H. Ohno, Electrochem. Commun., 2007, 9, 361–364 CrossRef CAS.
  156. V. V. Shumyantseva, T. V. Bulko, P. I. Koroleva, E. V. Shikh, A. A. Makhova, M. S. Kisel, I. V. Haidukevich and A. A. Gilep, Processes, 2022, 10, 383 CrossRef CAS.
  157. V. V. Shumyantseva, T. V. Bulko, G. P. Kuznetsova, N. F. Samenkova and A. I. Archakov, Biochem. Mosc., 2009, 74, 438–444 CrossRef CAS PubMed.
  158. S. Liu, L. Peng, X. Yang, Y. Wu and L. He, Anal. Biochem., 2008, 375, 209–216 CrossRef CAS PubMed.
  159. A. Fantuzzi, M. Fairhead and G. Gilardi, J. Am. Chem. Soc., 2004, 126, 5040–5041 CrossRef CAS PubMed.
  160. C. Garcia-Galan, Á. Berenguer-Murcia, R. Fernandez-Lafuente and R. C. Rodrigues, Adv. Synth. Catal., 2011, 353, 2885–2904 CrossRef CAS.
  161. C. M. Silveira, P. R. Rodrigues, W. Ghach, S. A. Pereira, F. Esteves, M. Kranendonk, M. Etienne and M. G. Almeida, ChemElectroChem, 2021, 8, 500–507 CrossRef CAS.
  162. T. Gul, R. Bischoff and H. P. Permentier, TrAC, Trends Anal. Chem., 2015, 70, 58–66 CrossRef CAS.
  163. D. P. Manica, Y. Mitsumori and A. G. Ewing, Anal. Chem., 2003, 75, 4572–4577 CrossRef CAS PubMed.
  164. P. J. R. Roche, S.-M. Ng, K. Page, N. Goddard and R. Narayanaswamy, Sens. Actuators B Chem., 2009, 139, 97–103 CrossRef CAS.
  165. Y. Lvov, R. Price, B. Gaber and I. Ichinose, Colloids Surf., A, 2002, 198–200, 375–382 CrossRef CAS.
  166. Z. Tang, Y. Wang, P. Podsiadlo and N. A. Kotov, Adv. Mater., 2006, 18, 3203–3224 CrossRef CAS.
  167. G.-X. Wang, W.-J. Bao, J. Wang, Q.-Q. Lu and X.-H. Xia, Electrochem. Commun., 2013, 35, 146–148 CrossRef.
  168. A. K. Yagati, J. Min and S. Cho, J. Electrochem. Soc., 2014, 161, G133–G140 CrossRef CAS.
  169. A. V. Kuzikov, R. A. Masamrekh, T. A. Filippova, Y. I. Haurychenka, A. A. Gilep, T. V. Shkel, N. V. Strushkevich, S. A. Usanov and V. V. Shumyantseva, Electrochim. Acta, 2020, 333, 135539 CrossRef CAS.
  170. M. A. Alonso-Lomillo, C. Yardimci, O. Domínguez-Renedo and M. J. Arcos-Martínez, Anal. Chim. Acta, 2009, 633, 51–56 CrossRef CAS PubMed.
  171. P. M. Gannett, J. Kabulski, F. A. Perez, Z. Liu, D. Lederman, C. W. Locuson, R. R. Ayscue, N. M. Thomsen and T. S. Tracy, J. Am. Chem. Soc., 2006, 128, 8374–8375 CrossRef CAS PubMed.
  172. L. A. Wollenberg, J. L. Kabulski, M. J. Powell, J. Chen, D. R. Flora, T. S. Tracy and P. M. Gannett, Appl. Biochem. Biotechnol., 2014, 172, 1293–1306 CrossRef CAS PubMed.
  173. A. Fantuzzi, L. H. Mak, E. Capria, V. Dodhia, P. Panicco, S. Collins and G. Gilardi, Anal. Chem., 2011, 83, 3831–3839 CrossRef CAS PubMed.
  174. D. Bartczak and A. G. Kanaras, Langmuir, 2011, 27, 10119–10123 CrossRef CAS PubMed.
  175. S. Keleştemur, M. Altunbek and M. Culha, Appl. Surf. Sci., 2017, 403, 455–463 CrossRef.
  176. J. Lu, H. Li, D. Cui, Y. Zhang and S. Liu, Anal. Chem., 2014, 86, 8003–8009 CrossRef CAS PubMed.
  177. L. Peng, X. Yang, Q. Zhang and S. Liu, Electroanalysis, 2008, 20, 803–807 CrossRef CAS.
  178. K. R. Jegannathan, S. Abang, D. Poncelet, E. S. Chan and P. Ravindra, Crit. Rev. Biotechnol., 2008, 28, 253–264 CrossRef CAS PubMed.
  179. J. F. Rusling and A. E. F. Nassar, J. Am. Chem. Soc., 1993, 115, 11891–11897 CrossRef CAS.
  180. J. F. Rusling, Acc. Chem. Res., 1998, 31, 363–369 CrossRef CAS.
  181. P. M. Guto and J. F. Rusling, Electrochem. Commun., 2006, 8, 455–459 CrossRef CAS.
  182. J. K. Cullison, F. M. Hawkridge, N. Nakashima and S. Yoshikawa, Langmuir, 1994, 10, 877–882 CrossRef CAS.
  183. J. D. Burgess, M. C. Rhoten and F. M. Hawkridge, Langmuir, 1998, 14, 2467–2475 CrossRef CAS.
  184. P. Quantin, E. Colaço, K. El Kirat, C. Egles, H. Ficheux and J. Landoulsi, ACS Omega, 2018, 3, 12535–12544 CrossRef CAS PubMed.
  185. F. Hua, T. Cui and Y. M. Lvov, Nano Lett., 2004, 4, 823–825 CrossRef CAS.
  186. Y. M. Lvov, Handbook of Surfaces and Interfaces of Materials, Elsevier, 2001, pp. 169–188 Search PubMed.
  187. Y. M. Lvov, Z. Lu, J. B. Schenkman, X. Zu and J. F. Rusling, J. Am. Chem. Soc., 1998, 120, 4073–4080 CrossRef CAS.
  188. J. Alvarez-Malmagro, A. R. Oliveira, C. Gutiérrez-Sánchez, B. Villajos, I. A. C. Pereira, M. Vélez, M. Pita and A. L. De Lacey, ACS Appl. Mater. Interfaces, 2021, 13, 11891–11900 CrossRef CAS PubMed.
  189. Q. Dai, L. Yang, Y. Wang, X. Cao, C. Yao and X. Xu, Anal. Bioanal. Chem., 2020, 412, 4703–4712 CrossRef CAS PubMed.
  190. M. Ray, S. D. Mhaske, S. K. Haram and S. Mazumdar, Anal. Biochem., 2021, 626, 114204 CrossRef CAS PubMed.
  191. V. E. V. Ferrero, L. Andolfi, G. Di Nardo, S. J. Sadeghi, A. Fantuzzi, S. Cannistraro and G. Gilardi, Anal. Chem., 2008, 80, 8438–8446 CrossRef CAS PubMed.
  192. R. Frank, C. Prönnecke, R. Azendorf, H.-G. Jahnke, A. G. Beck-Sickinger and A. A. Robitzki, Lab Chip, 2020, 20, 1449–1460 RSC.
  193. A. V. Kuzikov, T. V. Bulko, P. I. Koroleva, R. A. Masamrekh, S. S. Babkina, A. A. Gilep and V. V. Shumyantseva, Biochem. Mosc. Suppl. Ser. B Biomed. Chem., 2020, 14, 252–259 CrossRef.
  194. X. Xu, W. Wei, M. Huang, L. Yao and S. Liu, Chem. Commun., 2012, 48, 7802 RSC.
  195. N. Sultana, J. B. Schenkman and J. F. Rusling, Electroanalysis, 2007, 19, 2499–2506 CrossRef CAS.
  196. A. J. Bard and L. R. Faulkner, Electrochemical methods: fundamentals and applications, Wiley, New York, 2nd edn, 2001 Search PubMed.
  197. C. Baj-Rossi, G. D. Micheli and S. Carrara, Sensors, 2012, 12, 6520–6537 CrossRef CAS PubMed.
  198. T. Yuan, H. Permentier and R. Bischoff, TrAC, Trends Anal. Chem., 2015, 70, 50–57 CrossRef CAS.
  199. Q. Xue, D. Kato, T. Kamata, Q. Guo, T. You and O. Niwa, The Analyst, 2013, 138, 6463 RSC.
  200. V. V. Shumyantseva, S. Carrara, V. Bavastrello, D. Jason Riley, T. V. Bulko, K. G. Skryabin, A. I. Archakov and C. Nicolini, Biosens. Bioelectron., 2005, 21, 217–222 CrossRef CAS PubMed.
  201. H. F. Ozbakir and D. A. Sambade, J. Biosens. Bioelectron., 2015, 7, 1000193 Search PubMed.
  202. A. A. Makhova, V. V. Shumyantseva, E. V. Shich, T. V. Bulko, V. G. Kukes, O. S. Sizova, G. V. Ramenskaya, S. A. Usanov and A. I. Archakov, Bionanoscience, 2011, 1, 46–52 CrossRef.
  203. Y. Mie, M. Suzuki and Y. Komatsu, J. Am. Chem. Soc., 2009, 131, 6646–6647 CrossRef CAS PubMed.
  204. V. V. Shumyantseva, T. V. Bulko, Yu. O. Rudakov, G. P. Kuznetsova, N. F. Samenkova, A. V. Lisitsa, I. I. Karuzina and A. I. Archakov, J. Inorg. Biochem., 2007, 101, 859–865 CrossRef CAS PubMed.
  205. A. Shukla, E. M. Gillam, D. J. Mitchell and P. V. Bernhardt, Electrochem. Commun., 2005, 7, 437–442 CrossRef CAS.
  206. S. Y. Rhieu, D. R. Ludwig, V. S. Siu and G. T. R. Palmore, Electrochem. Commun., 2009, 11, 1857–1860 CrossRef CAS.
  207. D. P. Wasalathanthri, S. Malla, R. C. Faria and J. F. Rusling, Electroanalysis, 2012, 24, 2049–2052 CrossRef CAS PubMed.
  208. D. A. Buttry and F. C. Anson, J. Electroanal. Chem. Interfacial Electrochem., 1981, 130, 333–338 CrossRef CAS.
  209. S. S. Reginald, H. Lee, N. Fazil, B. Sharif, M. Lee, M. J. Kim, H. Beyenal and I. S. Chang, Commun. Biol., 2022, 5, 390 CrossRef CAS PubMed.
  210. C. C. Page, C. C. Moser, X. Chen and P. L. Dutton, Nature, 1999, 402, 47–52 CrossRef CAS PubMed.
  211. N. D. J. Yates, M. A. Fascione and A. Parkin, Chem. – Eur. J., 2018, 24, 12164–12182 CrossRef CAS PubMed.
  212. H. G. Rennke and M. A. Venkatachalam, J. Histochem. Cytochem., 1979, 27, 1352–1353 CrossRef CAS PubMed.
  213. M. Hoarau, S. Badieyan and E. N. G. Marsh, Org. Biomol. Chem., 2017, 15, 9539–9551 RSC.
  214. X. Yan, J. Tang, D. Tanner, J. Ulstrup and X. Xiao, Catalysts, 2020, 10, 1458 CrossRef CAS.
  215. K. P. Fears, B. Sivaraman, G. L. Powell, Y. Wu and R. A. Latour, Langmuir, 2009, 25, 9319–9327 CrossRef CAS PubMed.
  216. Y. Liu, T. L. Ogorzalek, P. Yang, M. M. Schroeder, E. N. G. Marsh and Z. Chen, J. Am. Chem. Soc., 2013, 135, 12660–12669 CrossRef CAS PubMed.
  217. N. Lalaoui, P. Rousselot-Pailley, V. Robert, Y. Mekmouche, R. Villalonga, M. Holzinger, S. Cosnier, T. Tron and A. Le Goff, ACS Catal., 2016, 6, 1894–1900 CrossRef CAS.
  218. W.-H. Shao, X.-E. Zhang, H. Liu, Z.-P. Zhang and A. E. G. Cass, Bioconjug. Chem., 2000, 11, 822–826 CrossRef CAS PubMed.
  219. A. Ménard, Y. Huang, P. Karam, G. Cosa and K. Auclair, Bioconjug. Chem., 2012, 23, 826–836 CrossRef PubMed.
  220. J. Park, S. Lee, Y. Kim and T. H. Yoo, Bioorg. Med. Chem., 2021, 30, 115946 CrossRef CAS PubMed.
  221. K. Yoshioka, D. Kato, T. Kamata and O. Niwa, Anal. Chem., 2013, 85, 9996–9999 CrossRef CAS PubMed.
  222. D. L. Johnson and L. L. Martin, J. Am. Chem. Soc., 2005, 127, 2018–2019 CrossRef CAS PubMed.
  223. M. Ferreira, S. Güney, I. Kuźniarska-Biernacka, O. S. G. P. Soares, J. L. Figueiredo, M. F. R. Pereira, I. C. Neves, A. M. Fonseca and P. Parpot, New J. Chem., 2021, 45, 12622–12633 RSC.
  224. R. Nerimetla, C. Walgama, V. Singh, S. D. Hartson and S. Krishnan, ACS Catal., 2017, 7, 3446–3453 CrossRef CAS.
  225. Y. Zhang, A. Iqbal, J. Zai, S.-Y. Zhang, H. Guo, X. Liu, I. Ul Islam, H. Fazal and X. Qian, Org. Chem. Front., 2022, 9, 436–444 RSC.
  226. C. Estavillo, Z. Lu, I. Jansson, J. B. Schenkman and J. F. Rusling, Biophys. Chem., 2003, 104, 291–296 CrossRef CAS PubMed.
  227. L. Jeftic and R. N. Adams, J. Am. Chem. Soc., 1970, 92, 1332–1337 CrossRef CAS PubMed.
  228. L. Mi, F. He, L. Jiang, L. Shangguan, X. Zhang, T. Ding, A. Liu, Y. Zhang and S. Liu, J. Electroanal. Chem., 2017, 804, 23–28 CrossRef CAS.
  229. C. Lütke Eversloh, M. Schulz, M. Wagner and T. A. Ternes, Water Res., 2015, 72, 293–304 CrossRef PubMed.
  230. D. Cui, L. Mi, X. Xu, J. Lu, J. Qian and S. Liu, Langmuir, 2014, 30, 11833–11840 CrossRef CAS PubMed.
  231. K. Elouarzaki, D. Cheng, A. C. Fisher and J.-M. Lee, Nat. Energy, 2018, 3, 574–581 CrossRef.
  232. E. A. Della Pia, J. E. Macdonald, M. Elliott and D. D. Jones, Small, 2012, 8, 2341–2344 CrossRef CAS PubMed.
  233. F. A. Al-Lolage, M. Meneghello, S. Ma, R. Ludwig and P. N. Bartlett, ChemElectroChem, 2017, 4, 1528–1534 CrossRef CAS.
  234. S. Zernia, R. Frank, R. H.-J. Weiße, H.-G. Jahnke, K. Bellmann-Sickert, A. Prager, B. Abel, N. Sträter, A. Robitzki and A. G. Beck-Sickinger, ChemCatChem, 2018, 10, 525–530 CrossRef CAS.
  235. J.-A. He, L. Samuelson, L. Li, J. Kumar and S. K. Tripathy, Langmuir, 1998, 14, 1674–1679 CrossRef CAS.
  236. J. Hodak, R. Etchenique, E. J. Calvo, K. Singhal and P. N. Bartlett, Langmuir, 1997, 13, 2708–2716 CrossRef CAS.
  237. X. Yu, G. A. Sotzing, F. Papadimitrakopoulos and J. F. Rusling, Anal. Chem., 2003, 75, 4565–4571 CrossRef CAS PubMed.
  238. J. Easterbrook, C. Lu, Y. Sakai and A. P. Li, Drug Metab. Dispos., 2001, 29, 141–144 CAS.
  239. K. Ryu and J. S. Dordick, Biochemistry, 1992, 31, 2588–2598 CrossRef CAS PubMed.
  240. T. T. Herskovits, B. Gadegbeku and H. Jaillet, J. Biol. Chem., 1970, 245, 2588–2598 CrossRef CAS PubMed.
  241. Y. Zhang, P. He and N. Hu, Electrochim. Acta, 2004, 49, 1981–1988 CrossRef CAS.
  242. W. E. Blumberg, J. Peisach, B. A. Wittenberg and J. B. Wittenberg, J. Biol. Chem., 1968, 243, 1854–1862 CrossRef CAS PubMed.
  243. A. M. Azevedo, D. M. F. Prazeres, J. M. S. Cabral and L. P. Fonseca, J. Mol. Catal. B: Enzym., 2001, 15, 147–153 CrossRef CAS.
  244. P. Singh, R. Prakash and K. Shah, Talanta, 2012, 97, 204–210 CrossRef CAS PubMed.
  245. T. Seng Wong, F. H. Arnold and U. Schwaneberg, Biotechnol. Bioeng., 2004, 85, 351–358 CrossRef PubMed.
  246. S. Sakalli, V. Burkina, V. Zlabek and G. Zamaratskaia, Toxicol. Mech. Methods, 2015, 25, 501–506 CAS.
  247. Horseradish Peroxidase (HRP) Enzymes, https://www.sigmaaldrich.com/US/en/technical-documents/technical-article/protein-biology/enzyme-activity-assays/peroxidase-enzymes, (accessed January 3, 2023).
  248. K.-F. Aguey-Zinsou, P. V. Bernhardt, J. J. De Voss and K. E. Slessor, Chem. Commun., 2003, 418–419 RSC.
  249. B. D. Fleming, Y. Tian, S. G. Bell, L.-L. Wong, V. Urlacher and H. A. O. Hill, Eur. J. Biochem., 2003, 270, 4082–4088 CrossRef CAS PubMed.
  250. D. L. Johnson, B. C. Lewis, D. J. Elliot, J. O. Miners and L. L. Martin, Biochem. Pharmacol., 2005, 69, 1533–1541 CrossRef CAS PubMed.
  251. J. F. Rusling, Pure Appl. Chem., 2001, 73, 1895–1905 CrossRef CAS.
  252. J. F. Rusling and C. J. Campbell, in Encyclopedia of surface and colloid science, ed. A. T. Hubbard, Marcel Dekker, New York, 2002, vol. 1, pp. 1754–1770 Search PubMed.
  253. H. Y. Karasulu, Expert Opin. Drug Deliv., 2008, 5, 119–135 CrossRef CAS PubMed.
  254. A. Vaze and J. F. Rusling, Langmuir, 2006, 22, 10788–10795 CrossRef CAS PubMed.
  255. A. Vaze, M. Parizo and J. F. Rusling, Langmuir, 2004, 20, 10943–10948 CrossRef CAS PubMed.
  256. V. A. Tamhane, D. G. Dhaware, N. Khandelwal, A. P. Giri and V. Panchagnula, J. Colloid Interface Sci., 2012, 383, 177–183 CrossRef CAS PubMed.
  257. S. A. Raya, I. Mohd Saaid, A. Abbas Ahmed and A. Abubakar Umar, J. Pet. Explor. Prod. Technol., 2020, 10, 1711–1728 CrossRef CAS.
  258. N. Kumar and R. N. Goyal, Electroanalysis, 2018, 30, 892–900 CrossRef CAS.
  259. J. F. Rusling, in Encyclopedia of Electrochemistry, ed. A. J. Bard, Wiley, 1st edn, 2003 Search PubMed.
  260. J. F. Rusling, Pure Appl. Chem., 2001, 73, 1895–1905 CrossRef CAS.
  261. A. C. Onuoha, X. Zu and J. F. Rusling, J. Am. Chem. Soc., 1997, 119, 3979–3986 CrossRef CAS.
  262. G. N. Kamau, M. P. Guto, B. Munge, V. Panchagnula and J. F. Rusling, Langmuir, 2003, 19, 6976–6981 CrossRef CAS.
  263. P. M. Guto, C. V. Kumar and J. F. Rusling, J. Phys. Chem. B, 2007, 111, 9125–9131 CrossRef CAS PubMed.
  264. P. M. Guto, C. V. Kumar and J. F. Rusling, Langmuir, 2008, 24, 10365–10370 CrossRef CAS PubMed.
  265. P. M. Guto and J. F. Rusling, J. Phys. Chem. B, 2005, 109, 24457–24464 CrossRef CAS PubMed.
  266. W. Nam, Acc. Chem. Res., 2007, 40, 522–531 CrossRef CAS PubMed.
  267. M. Mukherjee and A. Dey, ACS Cent. Sci., 2019, 5, 671–682 CrossRef CAS PubMed.
  268. T. Kamachi, T. Kouno, W. Nam and K. Yoshizawa, J. Inorg. Biochem., 2006, 100, 751–754 CrossRef CAS PubMed.
  269. J. C. Price, E. W. Barr, T. E. Glass, C. Krebs and J. M. Bollinger, J. Am. Chem. Soc., 2003, 125, 13008–13009 CrossRef CAS PubMed.
  270. P. Hlavica, Eur. J. Biochem., 2004, 271, 4335–4360 CrossRef CAS PubMed.
  271. M. N. Podgorski, J. S. Harbort, J. H. Z. Lee, G. T. H. Nguyen, J. B. Bruning, W. A. Donald, P. V. Bernhardt, J. R. Harmer and S. G. Bell, ACS Catal., 2022, 12, 1614–1625 CrossRef CAS.
  272. J. J. D. Sacramento and D. P. Goldberg, Acc. Chem. Res., 2018, 51, 2641–2652 CrossRef CAS PubMed.
  273. J. Yang, Y. Qi, J. A. V. Blodgett and T. A. Wencewicz, J. Nat. Prod., 2022, 85, 47–55 CrossRef CAS PubMed.
  274. Y. Mie, E. Tateyama and Y. Komatsu, Electrochim. Acta, 2014, 115, 364–369 CrossRef CAS.
  275. S. D. Mhaske, M. Ray and S. Mazumdar, Inorganica Chim. Acta, 2010, 363, 2804–2811 CrossRef CAS.
  276. J. Tian, J. Wang, Y. Li, M. Huang and J. Lu, J. Electrochem. Soc., 2017, 164, H470–H476 CrossRef CAS.
  277. P. Panicco, S. Castrignanò, S. J. Sadeghi, G. D. Nardo and G. Gilardi, Bioelectrochemistry, 2021, 138, 107729 CrossRef CAS PubMed.
  278. Y. Mie, Y. Yasutake, H. Takayama and T. Tamura, J. Catal., 2020, 384, 30–36 CrossRef CAS.
  279. X. Xu, G. Bai, L. Song, Q. Zheng, Y. Yao, S. Liu and C. Yao, Electrochim. Acta, 2017, 258, 1365–1374 CrossRef CAS.
  280. L. Shangguan, Y. Zhao, L. Mi, L. Jiang and S. Liu, J. Electroanal. Chem., 2016, 772, 46–51 CrossRef CAS.
  281. R. Nerimetla and S. Krishnan, Chem. Commun., 2015, 51, 11681–11684 RSC.
  282. J. Rittle and M. T. Green, Science, 2010, 330, 933–937 CrossRef CAS PubMed.
  283. D. F. Lewis and Y. Ito, Expert Opin. Drug Metab. Toxicol., 2008, 4, 1181–1186 CrossRef CAS PubMed.
  284. M. Pickl, S. Kurakin, F. G. Cantú Reinhard, P. Schmid, A. Pöcheim, C. K. Winkler, W. Kroutil, S. P. de Visser and K. Faber, ACS Catal., 2019, 9, 565–577 CrossRef CAS PubMed.
  285. H. Park, G. Park, W. Jeon, J.-O. Ahn, Y.-H. Yang and K.-Y. Choi, Biotechnol. Adv., 2020, 40, 107504 CrossRef PubMed.
  286. X. Zhang, Y. Hu, W. Peng, C. Gao, Q. Xing, B. Wang and A. Li, Front. Chem., 2021, 9, 649000 CrossRef CAS PubMed.
  287. U. A. Argikar, J. C. Cloyd, A. K. Birnbaum, I. E. Leppik, J. Conway, S. Kshirsagar, W. S. Oetting, E. C. Klein and R. P. Remmel, Epilepsy Res., 2006, 71, 54–63 CrossRef CAS PubMed.
  288. D. Li, X. Huang, K. Han and C.-G. Zhan, J. Am. Chem. Soc., 2011, 133, 7416–7427 CrossRef CAS PubMed.
  289. R.-J. Li, Z. Zhang, C. G. Acevedo-Rocha, J. Zhao and A. Li, Green Synth. Catal., 2020, 1, 52–59 CrossRef.
  290. A. R. Jalalvand, Microchem. J., 2020, 157, 105104 CrossRef CAS.
  291. D. Holtmann, K.-M. Mangold and J. Schrader, Biotechnol. Lett., 2009, 31, 765–770 CrossRef CAS PubMed.
  292. V. V. Shumyantseva, T. V. Bulko, A. V. Kuzikov, R. A. Masamrekh, A. Y. Konyakhina, I. Romanenko, J. B. Max, M. Köhler, A. A. Gilep, S. A. Usanov, D. V. Pergushov, F. H. Schacher and L. V. Sigolaeva, Electrochim. Acta, 2020, 336, 135579 CrossRef CAS.
  293. D. P. Wasalathanthri, D. Li, D. Song, Z. Zheng, D. Choudhary, I. Jansson, X. Lu, J. B. Schenkman and J. F. Rusling, Chem. Sci., 2015, 6, 2457–2468 RSC.
  294. B. Song, S. Pan, C. Tang, D. Li and J. F. Rusling, Anal. Chem., 2013, 85, 11061–11067 CrossRef CAS PubMed.
  295. N. Kumar, Rosy and R. N. Goyal, Talanta, 2017, 166, 215–222 CrossRef CAS PubMed.
  296. R. N. T. Kankanamage, A. B. Ghosh, D. Jiang, K. Gkika, T. Keyes, L. A. Achola, S. Suib and J. F. Rusling, Chem. Res. Toxicol., 2020, 33, 2072–2086 Search PubMed.
  297. M. Kazui, Y. Nishiya, T. Ishizuka, K. Hagihara, N. A. Farid, O. Okazaki, T. Ikeda and A. Kurihara, Drug Metab. Dispos., 2010, 38, 92–99 CrossRef CAS PubMed.
  298. T. Coleman, A. M. Kirk, J. H. Z. Lee, D. Z. Doherty, J. B. Bruning, E. H. Krenske, J. J. De Voss and S. G. Bell, ACS Catal., 2022, 12, 1258–1267 CrossRef CAS.
  299. S. G. Bell, J.-A. Stevenson, H. D. Boyd, S. Campbell, A. D. Riddle, E. L. Orton and L.-L. Wong, Chem. Commun., 2002, 490–491 RSC.
  300. M. M. Chen, P. S. Coelho and F. H. Arnold, Adv. Synth. Catal., 2012, 354, 964–968 CrossRef CAS.
  301. P. Meinhold, M. W. Peters, M. M. Y. Chen, K. Takahashi and F. H. Arnold, ChemBioChem, 2005, 6, 1765–1768 CrossRef CAS PubMed.
  302. E. T. Farinas, U. Schwaneberg, A. Glieder and F. H. Arnold, Adv. Synth. Catal., 2001, 343, 601–606 CrossRef CAS.
  303. R. Karande, L. Debor, D. Salamanca, F. Bogdahn, K.-H. Engesser, K. Buehler and A. Schmid, Biotechnol. Bioeng., 2016, 113, 52–61 CrossRef CAS PubMed.
  304. A. Hoschek, J. Toepel, A. Hochkeppel, R. Karande, B. Bühler and A. Schmid, Biotechnol. J., 2019, 14, 1800724 CrossRef PubMed.
  305. T. Fukami, M. Katoh, H. Yamazaki, T. Yokoi and M. Nakajima, Chem. Res. Toxicol., 2008, 21, 720–725 Search PubMed.
  306. S. C. Dodani, J. K. B. Cahn, T. Heinisch, S. Brinkmann-Chen, J. A. McIntosh and F. H. Arnold, ChemBioChem, 2014, 15, 2259–2267 CrossRef CAS PubMed.
  307. P. Dydio, H. M. Key, H. Hayashi, D. S. Clark and J. F. Hartwig, J. Am. Chem. Soc., 2017, 139, 1750–1753 CrossRef CAS PubMed.
  308. I. S. Williams, L. Gatchie, S. B. Bharate and B. Chaudhuri, ACS Omega, 2018, 3, 8903–8912 CrossRef CAS PubMed.
  309. M. P. Mayhew, V. Reipa, M. J. Holden and V. L. Vilker, Biotechnol. Prog., 2000, 16, 610–616 CrossRef CAS PubMed.
  310. E. N. Jacobsen, W. Zhang and M. L. Guler, J. Am. Chem. Soc., 1991, 113, 6703–6704 CrossRef CAS.
  311. G. Yin, A. M. Danby, D. Kitko, J. D. Carter, W. M. Scheper and D. H. Busch, Inorg. Chem., 2007, 46, 2173–2180 CrossRef CAS PubMed.
  312. G. Yin, M. Buchalova, A. M. Danby, C. M. Perkins, D. Kitko, J. D. Carter, W. M. Scheper and D. H. Busch, Inorg. Chem., 2006, 45, 3467–3474 CrossRef CAS PubMed.
  313. H. F. Fisher, E. E. Conn, B. Vennesland and F. H. Westheimer, J. Biol. Chem., 1953, 202, 687–697 CrossRef CAS PubMed.
  314. J. T. Groves and R. S. Myers, J. Am. Chem. Soc., 1983, 105, 5791–5796 CrossRef CAS.
  315. V. Polic, K. J. Cheong, F. Hammerer and K. Auclair, Adv. Synth. Catal., 2017, 359, 3983–3989 CrossRef CAS.
  316. R. Lonsdale, J. N. Harvey and A. J. Mulholland, J. Phys. Chem. B, 2010, 114, 1156–1162 CrossRef CAS PubMed.
  317. P. Rydberg, R. Lonsdale, J. N. Harvey, A. J. Mulholland and L. Olsen, J. Mol. Graph. Model., 2014, 52, 30–35 CrossRef CAS PubMed.
  318. T. Coleman, A. M. Kirk, R. R. Chao, M. N. Podgorski, J. S. Harbort, L. R. Churchman, J. B. Bruning, P. V. Bernhardt, J. R. Harmer, E. H. Krenske, J. J. De Voss and S. G. Bell, ACS Catal., 2021, 11, 1995–2010 CrossRef CAS.
  319. C.-H. Yun, K.-H. Kim, D.-H. Kim, H.-C. Jung and J.-G. Pan, Trends Biotechnol., 2007, 25, 289–298 CrossRef CAS PubMed.
  320. L. Zhang and Q. Wang, ChemBioChem, 2022, 23, e202100439 CAS.
  321. D. Kumar, S. P. de Visser and S. Shaik, Chem. – Eur. J., 2005, 11, 2825–2835 CrossRef CAS PubMed.
  322. S. Jin, T. M. Makris, T. A. Bryson, S. G. Sligar and J. H. Dawson, J. Am. Chem. Soc., 2003, 125, 3406–3407 CrossRef CAS PubMed.
  323. R. R. Chao, I. C.-K. Lau, T. Coleman, L. R. Churchman, S. A. Child, J. H. Z. Lee, J. B. Bruning, J. J. De Voss and S. G. Bell, Chem. – Eur. J., 2021, 27, 14765–14777 CrossRef CAS PubMed.
  324. A. Çelik, D. Sperandio, R. E. Speight and N. J. Turner, Org. Biomol. Chem., 2005, 3, 2688 RSC.
  325. H. Kawana, T. Miwa, Y. Honda and T. Furuya, ACS Omega, 2022, 7, 20259–20266 CrossRef CAS PubMed.
  326. A. W. Munro, K. J. McLean, J. L. Grant and T. M. Makris, Biochem. Soc. Trans., 2018, 46, 183–196 CrossRef CAS PubMed.
  327. J. Zitko, B. Servusová, A. Janoutová, P. Paterová, J. Mandíková, V. Garaj, M. Vejsová, J. Marek and M. Doležal, Bioorg. Med. Chem., 2015, 23, 174–183 CrossRef CAS PubMed.
  328. U. Bussy and M. Boujtita, Chem. Res. Toxicol., 2014, 27, 1652–1668 Search PubMed.
  329. I. Geronimo, C. A. Denning, D. K. Heidary, E. C. Glazer and C. M. Payne, Biophys. J., 2018, 115, 1251–1263 CrossRef CAS PubMed.
  330. S. H. Lee, Y.-C. Kwon, D.-M. Kim and C. B. Park, Biotechnol. Bioeng., 2013, 110, 383–390 CrossRef CAS PubMed.
  331. P. Rydberg, U. Ryde and L. Olsen, J. Chem. Theory Comput., 2008, 4, 1369–1377 CrossRef CAS PubMed.
  332. L. T. Burka, F. P. Guengerich, R. J. Willard and T. L. Macdonald, J. Am. Chem. Soc., 1985, 107, 2549–2551 CrossRef CAS.
  333. C. Li, W. Wu, K. Cho and S. Shaik, Chem. – Eur. J., 2009, 15, 8492–8503 CrossRef CAS PubMed.
  334. F. P. Guengerich, C.-H. Yun and T. L. Macdonald, J. Biol. Chem., 1996, 271, 27321–27329 CrossRef CAS PubMed.
  335. T. Johansson, L. Weidolf and U. Jurva, Rapid Commun. Mass Spectrom., 2007, 21, 2323–2331 CrossRef CAS PubMed.
  336. E. Nouri-Nigjeh, H. P. Permentier, R. Bischoff and A. P. Bruins, Anal. Chem., 2010, 82, 7625–7633 CrossRef CAS PubMed.
  337. A. H. Meyer, A. Dybala-Defratyka, P. J. Alaimo, I. Geronimo, A. D. Sanchez, C. J. Cramer and M. Elsner, Dalton Trans., 2014, 43, 12175–12186 RSC.
  338. A. A. Najmi, R. Bischoff and H. P. Permentier, Molecules, 2022, 27, 3293 CrossRef CAS PubMed.
  339. F. Rao, Z. Chen, D. Zhou, Y. Kang, L. Guo and Y. Xue, J. Mol. Graph. Model., 2018, 84, 109–117 CrossRef CAS PubMed.
  340. A. M. Carmona-Ribeiro, T. Prieto and I. L. Nantes, Front. Mol. Biosci., 2015, 2, 50 Search PubMed.
  341. S. Lütz, E. Steckhan and A. Liese, Electrochem. Commun., 2004, 6, 583–587 CrossRef.
  342. C. Regalado, J. A. Asenjo and D. L. Pyle, Enzyme Microb. Technol., 1996, 18, 332–339 CrossRef CAS PubMed.
  343. A. Henriksen, Protein Sci., 2001, 10, 108–115 CrossRef CAS PubMed.
  344. Y. Bhandari, H. Sajwan, P. Pandita and V. Koteswara Rao, Biocatal. Biotransform., 2022, 1–18 CrossRef.
  345. M. Hofrichter, Enzyme Microb. Technol., 2002, 30, 454–466 CrossRef CAS.
  346. O. D. V. Biko, M. Viljoen-Bloom and W. H. van Zyl, Enzyme Microb. Technol., 2020, 141, 109669 CrossRef CAS PubMed.
  347. W. Zhang, W. Ye, Y. Wang and Y. Yan, Biotechnol. J., 2022, 17, 2200098 CrossRef CAS PubMed.
  348. S.-H. Cho, J. Shim, S.-H. Yun and S.-H. Moon, Appl. Catal. Gen., 2008, 337, 66–72 CrossRef CAS.
  349. K. S. Yang, G. Mul and J. A. Moulijn, Electrochim. Acta, 2007, 52, 6304–6309 CrossRef CAS.
  350. R. D. C. Soltani, A. Rezaee, A. R. Khataee and H. Godini, Res. Chem. Intermed., 2013, 39, 4277–4286 CrossRef CAS.
  351. V. Viswanathan, H. A. Hansen and J. K. Nørskov, J. Phys. Chem. Lett., 2015, 6, 4224–4228 CrossRef CAS PubMed.
  352. W. R. P. Barros, T. Ereno, A. C. Tavares and M. R. V. Lanza, ChemElectroChem, 2015, 2, 714–719 CrossRef CAS.
  353. A. Khataee, S. Sajjadi, S. Rahim Pouran and A. Hasanzadeh, J. Ind. Eng. Chem., 2017, 56, 312–320 CrossRef CAS.
  354. Y.-C. Yin, Y. Shi, Y.-B. Sun, D.-R. Yang, Y. Liu and X.-H. Xia, ACS Appl. Energy Mater., 2021, 4, 10843–10848 CrossRef CAS.
  355. K. B. Lee, M. B. Gu and S.-H. Moon, J. Chem. Technol. Biotechnol., 2001, 76, 811–819 CrossRef CAS.
  356. H. Tian, S. Mu, H. Li, X. Wu and Z. Lu, ChemCatChem, 2012, 4, 1850–1855 CrossRef CAS.
  357. K. Lee and S.-H. Moon, J. Biotechnol., 2003, 102, 261–268 CrossRef CAS PubMed.
  358. S. Lütz, K. Vuorilehto and A. Liese, Biotechnol. Bioeng., 2007, 98, 525–534 CrossRef PubMed.
  359. C. Kohlmann and S. Lütz, Eng. Life Sci., 2006, 6, 170–174 CrossRef CAS.
  360. T. Krieg, S. Hüttmann, K.-M. Mangold, J. Schrader and D. Holtmann, Green Chem., 2011, 13, 2686 RSC.
  361. M. Xia, Z. Wang, L. Du, Z. Fu, G. Jiang, D. Lu, J. Wu and Z. Liu, ChemCatChem, 2022, 14, e202101732 CrossRef CAS.
  362. A. E. W. Horst, S. Bormann, J. Meyer, M. Steinhagen, R. Ludwig, A. Drews, M. Ansorge-Schumacher and D. Holtmann, J. Mol. Catal. B: Enzym., 2016, 133, S137–S142 CrossRef.
  363. S. Bormann, M. M. C. H. Schie, T. P. De Almeida, W. Zhang, M. Stöckl, R. Ulber, F. Hollmann and D. Holtmann, ChemSusChem, 2019, 12, 4759–4763 CrossRef CAS PubMed.
  364. X. Zhu, Y. Ding, S. Li, Y. Chen, Q. Zhai and Y. Jiang, ACS Sustainable Chem. Eng., 2022, 10, 12497–12503 CrossRef CAS.
  365. N. Ž. Šekuljica, J. R. Jovanović, S. M. Jakovetić Tanasković, N. D. Ognjanović, I. V. Gazikalović, Z. D. Knežević-Jugović and D. Ž. Mijin, Biotechnol. Prog., 2020, 36, e2991 CrossRef PubMed.
  366. B. E. Keshta, A. H. Gemeay and A. A. Khamis, Environ. Sci. Pollut. Res., 2022, 29, 6633–6645 CrossRef CAS PubMed.
  367. J. Wang, Z. Shu, Z. Chen, J. Su and C. Liu, J. Clean. Prod., 2022, 337, 130489 CrossRef CAS.
  368. D. Tarasek, B. Gąsowska-Bajger, B. Frąckowiak-Wojtasek, C. Kersten, M. Jewgiński, Ł. Kołodziej, R. Latajka and H. Wojtasek, J. Mol. Struct., 2022, 1252, 132169 CrossRef CAS.
  369. G. Zambrano, A. Sekretareva, D. D’Alonzo, L. Leone, V. Pavone, A. Lombardi and F. Nastri, RSC Adv., 2022, 12, 12947–12956 RSC.
  370. J. Rocha-Martin, S. Velasco-Lozano, J. M. Guisán and F. López-Gallego, Green Chem., 2014, 16, 303–311 RSC.
  371. V. M. Samokyszyn, J. P. Freeman, K. R. Maddipati and R. V. Lloyd, Chem. Res. Toxicol., 1995, 8, 349–355 Search PubMed.
  372. H. Nunavath, C. Banoth, V. R. Talluri and B. Bhukya, Bioinformation, 2016, 12, 318–323 Search PubMed.
  373. J. Xu, T. Tang, K. Zhang, S. Ai and H. Du, Process Biochem., 2011, 46, 1160–1165 CrossRef CAS.
  374. Z. Lu, W. Hou, X. Wu, H. Li and J. Chen, Chem. Lett., 2007, 36, 564–565 CrossRef CAS.
  375. K. Piontek, A. T. Smith and W. Blodig, Biochem. Soc. Trans., 2001, 29, 111–116 CrossRef CAS PubMed.
  376. S. Parveen, M. Asgher and M. Bilal, Environ. Technol. Innov., 2021, 21, 101226 CrossRef CAS.
  377. V. D. Giap, D. H. Nghi, L. H. Cuong and D. T. Quynh, BioResources, 2022, 17, 4480–4498 CAS.
  378. E. E. Ferapontova, J. Castillo and L. Gorton, Biochim. Biophys. Acta, Gen. Subj., 2006, 1760, 1343–1354 CrossRef CAS PubMed.
  379. C. Yang, S. Maldonado and C. R. J. Stephenson, ACS Catal., 2021, 11, 10104–10114 CrossRef CAS.
  380. Y. Ding, K. Cui, X. Liu, Q. Xie, Z. Guo and Y. Chen, J. Hazard. Mater., 2022, 431, 128544 CrossRef CAS PubMed.
  381. Z. Liu, Y. Lv, A. Zhu and Z. An, ACS Macro Lett., 2018, 7, 1–6 CrossRef PubMed.
  382. E. Gantumur, S. Sakai, M. Nakahata and M. Taya, ACS Macro Lett., 2017, 6, 485–488 CrossRef CAS PubMed.
  383. A. Bindoli, J. M. Fukuto and H. J. Forman, Antioxid. Redox Signal., 2008, 10, 1549–1564 CrossRef CAS PubMed.
  384. V. M. Dembitsky and M. Srebnik, Eurasian Chem.-Technol. J., 2017, 4, 221 CrossRef.
  385. E. Baciocchi, M. F. Gerini, P. J. Harvey, O. Lanzalunga and S. Mancinelli, Eur. J. Biochem., 2000, 267, 2705–2710 CrossRef CAS PubMed.
  386. S. Sato, Y. Ohashi, M. Kojima, T. Watanabe, Y. Honda and T. Watanabe, Chemosphere, 2009, 77, 798–804 CrossRef CAS PubMed.
  387. H. Tang, Y. Zhang, Y. Deng, S. Du, D. Li, Z. Wang, H. Li, X. Gao and F. Wang, Catal. Lett., 2022, 152, 720–731 CrossRef CAS.
  388. H. Rabiee, L. Ge, X. Zhang, S. Hu, M. Li and Z. Yuan, Energy Environ. Sci., 2021, 14, 1959–2008 RSC.
  389. J. J. Mieyal and D. W. Starke, Methods in Enzymology, Elsevier, 1994, vol. 231, pp. 573–598 Search PubMed.
  390. S. Aoun and M. Baboulène, J. Mol. Catal. B: Enzym., 1998, 4, 101–109 CrossRef CAS.
  391. B.-A. Kaup, U. Piantini, M. Wüst and J. Schrader, Appl. Microbiol. Biotechnol., 2007, 73, 1087–1096 CrossRef CAS PubMed.
  392. M. Hobisch, D. Holtmann, P. Gomez de Santos, M. Alcalde, F. Hollmann and S. Kara, Biotechnol. Adv., 2021, 51, 107615 CrossRef CAS PubMed.
  393. M. Hofrichter, H. Kellner, R. Herzog, A. Karich, C. Liers, K. Scheibner, V. W. Kimani and R. Ullrich, in Grand Challenges in Fungal Biotechnology, ed. H. Nevalainen, Springer International Publishing, Cham, 2020, pp. 369–403 Search PubMed.
  394. D. Holtmann, T. Krieg, L. Getrey and J. Schrader, Catal. Commun., 2014, 51, 82–85 CrossRef CAS.
  395. C. Laane, A. Weyland and M. Franssen, Enzyme Microb. Technol., 1986, 8, 345–348 CrossRef CAS.
  396. F. Arnold, Trends Biotechnol., 1990, 8, 244–249 CrossRef CAS PubMed.
  397. A. Shaw, Curr. Opin. Struct. Biol., 1996, 6, 546–550 CrossRef CAS PubMed.
  398. P. S. Coelho, E. M. Brustad, A. Kannan and F. H. Arnold, Science, 2013, 339, 307–310 CrossRef CAS PubMed.
  399. S. V. Athavale, S. Gao, A. Das, S. C. Mallojjala, E. Alfonzo, Y. Long, J. S. Hirschi and F. H. Arnold, J. Am. Chem. Soc., 2022, 144, 19097–19105 CrossRef CAS PubMed.
  400. T. Cai, H. Sun, J. Qiao, L. Zhu, F. Zhang, J. Zhang, Z. Tang, X. Wei, J. Yang, Q. Yuan, W. Wang, X. Yang, H. Chu, Q. Wang, C. You, H. Ma, Y. Sun, Y. Li, C. Li, H. Jiang, Q. Wang and Y. Ma, Science, 2021, 373, 1523–1527 CrossRef CAS PubMed.
  401. T. Zheng, M. Zhang, L. Wu, S. Guo, X. Liu, J. Zhao, W. Xue, J. Li, C. Liu, X. Li, Q. Jiang, J. Bao, J. Zeng, T. Yu and C. Xia, Nat. Catal., 2022, 5, 388–396 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2023