Skip to content
Publicly Available Published by De Gruyter September 4, 2017

Strategies for the construction of insect P450 fusion enzymes

  • Lea Talmann , Jochen Wiesner and Andreas Vilcinskas EMAIL logo

Abstract

Cytochrome P450 monooxygenases (P450s) are ubiquitous enzymes with a broad substrate spectrum. Insect P450s are known to catalyze reactions such as the detoxification of insecticides and the synthesis of hydrocarbons, which makes them useful for many industrial processes. Unfortunately, it is difficult to utilize P450s effectively because they must be paired with cytochrome P450 reductases (CPRs) to facilitate electron transfer from reduced nicotinamide adenine dinucleotide phosphate (NADPH). Furthermore, eukaryotic P450s and CPRs are membrane-anchored proteins, which means they are insoluble and therefore difficult to purify when expressed in their native state. Both challenges can be addressed by creating fusion proteins that combine the P450 and CPR functions while eliminating membrane anchors, allowing the production and purification of soluble multifunctional polypeptides suitable for industrial applications. Here we discuss several strategies for the construction of fusion enzymes combining insect P450 with CPRs.

1 Introduction

The cytochrome P450 monooxygenases (P450s) are a ubiquitous superfamily of heme-thiolated monooxygenases that have been strongly conserved during evolution [1]. P450s fulfil diverse functions, including the metabolism of xenobiotics and drugs [2], steroid biosynthesis [3] and the assimilation of carbon sources for growth [4]. They were first discovered in 1958 and were isolated from mammalian liver microsomes in 1964 [5], [6], [7]. More than 1000 P450 families have been described and ~21,000 different P450s are known, with the number still growing [8]. A P450 nomenclature was introduced in 1987 comprising the prefix CYP followed by a family number, a subfamily letter and finally a number representing the specific enzyme, e.g. CYP6A1 [9]. All P450s form similar secondary and tertiary structures [10], [11], [12] which combine with a heme cofactor that forms the catalytic reaction center. Heme comprises a central iron atom bound to a protoporphyrin ring by four nitrogen ligands forming a planar structure. The iron atom is additionally coordinated by an evolutionarily conserved axial cysteine ligand (FXXGXXXCXG) located near the C-terminus of the P450 [1]. The iron atom is often associated with water as a sixth ligand. The heme cofactor is responsible for Soret band absorbance at 420 nm when it is coordinated with water. The absorbance signal shows a typical shift to 450 nm when carbon monoxide is bound at the opposite axial position, which is the basis of the superfamily name: pigment 450 [6].

1.1 Reactions catalyzed by cytochrome P450 monooxygenases

P450s cannot catalyze reactions on their own because they need to accept electrons from a cytochrome P450 reductase (CPR) redox partner, as shown in Figure 1. The reaction requires two electrons, derived from a nicotinamide cofactor [13]. Prosthetic groups, such as flavin adenine dinucleotide (FAD) and flavin mononucleotide (FMN) derived from riboflavin, serve as electron transfer centers. P450s activate molecular oxygen and catalyze unique one-step C–H bond oxidations by the insertion of one oxygen atom into the substrate while the other forms water [14], [15]. They can catalyze diverse reactions (Figure 2) and the nature of these reactions and the corresponding substrates does not appear to depend on the sequence of the P450 or its evolutionary proximity to other P450s [1]. Insects and mammals produce approximately 80 different P450s but in plants there is much greater diversity, e.g. there are 286 P450s in Arabidopsis thaliana. Plant P450s are designated as A-type or non-A-type. The former are specific to plants and fulfil functions such as herbicide metabolism and the synthesis of volatiles involved in plant–insect interactions, whereas the latter are more like the P450s in other phyla and fulfil similar roles, e.g. sterol and fatty acid metabolism [17]. The most common reaction is monooxygenation, resulting in the hydroxylation or epoxidation of a carbon center (Figure 2A). Others include oxidation and dealkylation at a heteroatom, and complex reactions such as single electron reductions, desaturations and ring modifications alongside typical oxidations [18] (Figure 2B). These reactions are often regiospecific and stereospecific, and thus difficult to replicate by total chemical synthesis. The selectivity of P450s is one reason they are valued in the context of industrial processes.

Figure 1: Schematic representation of the P450-redox partner interaction. P450, cytochrome P450 monooxygenase; CPR, cytochrome P450 reductase; FAD, flavin adenine dinucleotide; FMN, flavin mononucleotide; NADPH, reduced nicotinamide adenine dinucleotide phosphate.
Figure 1:

Schematic representation of the P450-redox partner interaction. P450, cytochrome P450 monooxygenase; CPR, cytochrome P450 reductase; FAD, flavin adenine dinucleotide; FMN, flavin mononucleotide; NADPH, reduced nicotinamide adenine dinucleotide phosphate.

Figure 2: (A) Common and (B) uncommon reactions catalyzed by cytochrome P450 monooxygenases [16].
Figure 2:

(A) Common and (B) uncommon reactions catalyzed by cytochrome P450 monooxygenases [16].

1.2 Insect P450s and their potential applications

Insect P450s are structurally and functionally similar to their better-characterized mammalian counterparts. More than 100 enzymes have been identified from diverse insect species, and many are components of important metabolic pathways [19]. They facilitate growth, development, feeding [1] and the degradation of endogenous compounds and xenobiotics [20]. An important example of the role of insect P450s in development is provided by the so-called Halloween genes: phantom (CYP306A1), disembodied (CYP302A1), shadow (CYP315A1), and shade (CYP314A1). These are responsible for the four last hydroxylation steps that convert steroid precursors into ecdysteroids, with CYP314A1 catalyzing the final conversion of ecdysterone to its more active derivative 20-hydroxyecdystone [21]. The importance of insect P450s in xenobiotic degradation is exemplified by their role in resistance to plant toxins and pesticides, which is often achieved by the overexpression of P450 genes targeting these compounds [1], [22]. Insect P450s can be microsomal or mitochondrial. Microsomal P450s depend on a CPR or cytochrome b5 (or both) as an electron donor, whereas mitochondrial P450s depend on an adrenodoxin-like ferredoxin coupled to adrenodoxin reductase. Six P450 families are known in insects, five of which are not found in other organisms (CYP6, CYP9, CYP12, CYP18 and CYP28). The expression levels of P450s in insects can vary across different life stages [23] and they are expressed in several tissues, such as the Malpighian tubules, fat body and midgut [24]. There is no clear phylogenetic distinction between the P450s involved in physiological homeostasis and those required for detoxification, suggesting that the function of P450s can change over evolutionary timescales [24].

The first purified insect CPR was prepared from the common housefly (Musca domestica) and was shown to be similar to mammalian CPRs [25]. The first purified insect P450 was the xenobiotic-degrading enzyme CYP6A1 also from the housefly, which was reconstituted in an Escherichia coli system allowing the identification of aldrin and heptachlor as substrates [26]. CYP6A1 has a spacious heme active site which is compatible with several substrate geometries and orientations [27]. Similarly, CYP6G1 from the fruit fly (Drosophila melanogaster) can be classed as a multi-pesticide degrading enzyme because it confers resistance towards several structurally unrelated compounds [28]. Insects are known to produce hydrocarbons that prevent desiccation and that also function as contact pheromones [29]. The underlying biosynthesis pathway involves the decarbonylation of aldehydes to form alkanes by CO2 cleavage. The first P450 shown to catalyze this unusual reaction was D. melanogaster CYP4G1 [30].

The large number of different P450s produced by insect species combined with the diversity of insects offers a vast library of enzymes with a broad range of catalytic activities. These could be used to investigate the origin of pesticide resistance and to facilitate the development of new insecticides that inhibit P450s [24]. Furthermore, insect P450s could also be deployed as industrial enzymes for the production of fine chemicals and pharmaceuticals that cannot be produced economically by total chemical synthesis [31] or for the development of new processes based on bioelectrocatalysis [32]. Finally, they could be used for remediation purposes, such as the removal of pesticide residues and other xenobiotics from wastewater and the environment [33].

1.3 Challenges hindering the production of recombinant P450s

The functional analysis and exploitation of insect P450s is challenging because they are membrane-bound enzymes that require various electron donors [1]. The standard approach to determine the catalytic activity of P450s is heterologous expression and reconstitution [22], but the enzymes must be reconstituted in a system containing detergents and phospholipids for solubilization, and the correct CPR components and cofactors must be supplied [34], [35]. P450s do not tolerate solvents well, so many efforts to solubilize the enzymes render them inactive [31]. Because P450s have diverse substrates, the activity of a recombinant enzyme can only be established by screening a library of substrates, some of which are also only sparing soluble [31]. Finally, the industrial application of P450s is also hampered by the need for the expensive electron donor nicotinamide adenine dinucleotide phosphate (NADPH) [31], [36], [37]. Catalytically active insect P450s have been expressed in E. coli [38] and insect cell lines [30] but in all cases the yields were insufficient for industrial application.

1.4 Addressing the challenges by constructing P450 fusion enzymes

The challenges described above can potentially be overcome by constructing P450 fusion enzymes that incorporate the necessary CPR components and eliminate sections of the polypeptide that hinder expression and reconstitution, such as the membrane anchor. Nature has provided promising evidence to support this approach because several natural P450 fusion enzymes have already been discovered [39] as listed in Table 1. CYP102A1 and CYP11B2 have been studied in detail to develop strategies for the design of artificial fusion proteins and protein evolution methods, and these are discussed below.

Table 1:

Examples of natural P450 fusion enzymes.

NameOriginDomains (N→C)Reference
CYP102A1/BM3Bacillus megateriumP450-FMN-FAD[40]
CYP11B2/RhFRhodococcus sp. strain NCIMB 9784P450-FMN-Fe/S Red[41]
CYP51fxMethylococcus capsulatusP450-Fdx[42]
Xp1ARhodococcus rhodochrousFMN-P450[43]
CYP221A1Pseudomonas fluorescensAcyl CoS-DeH-P450[44]
CYP5253A1MimivirusP450-?[45]
CYP55A1Fusarium oxysporumP450[39]
  1. The domain structure is shown from the N-terminus. P450, cytochrome P450 monooxygenase; FMN, flavin mononucleotide; FAD, flavin adenine dinucleotide; FMN-Fe/S Red, FMN-containing reductase with a [2Fe2S] ferredoxin-like center; Fdx, ferredoxin domain; Acyl CoS-DeH, P450-acyl-CoA dehydrogenase; ?, protein of unknown function containing several putative post-translational modification sites.

1.4.1 CYP102A1 (BM3)

Bacillus megaterium CYP102A1 (also known as BM3) is a fatty acid hydroxylase which catalyzes reactions without the assistance of additional proteins. It is a natural fusion protein, in which the N-terminal catalytic P450 domain is covalently attached via a linker to the C-terminal redox domain, constituting an entire class II P450 system [46] typical of microsomal eukaryotic P450s [47]. As shown in Figure 3, the 66 kDa redox domain (B. megaterium reductase, BMR) binds FAD and FMN as prosthetic groups, making it functionally similar to eukaryotic CPRs [46]. Indeed, BMR shares ~33% sequence identity with mammalian hepatic CPR. The 55 kDa P450 domain shares ~25% sequence identity with fatty acid ω-hydroxylases of the CYP4 family [48].

Figure 3: Schematic representation of the natural P450-redox partner fusion enzyme BM3. BMR, Bacillus megaterium reductase domain; P450, cytochrome P450 monooxygenase domain; FAD, flavin adenine dinucleotide; FMN, flavin mononucleotide; NADP+ and NADPH, oxidized and reduced nicotinamide adenine dinucleotide phosphate, respectively.
Figure 3:

Schematic representation of the natural P450-redox partner fusion enzyme BM3. BMR, Bacillus megaterium reductase domain; P450, cytochrome P450 monooxygenase domain; FAD, flavin adenine dinucleotide; FMN, flavin mononucleotide; NADP+ and NADPH, oxidized and reduced nicotinamide adenine dinucleotide phosphate, respectively.

Electron transfer from the BMR domain to the heme group influences the enzyme activity. The linker length is more important for enzyme activity than the amino acid composition because it determines the correct relative positions of each domain [49]. BM3 has the highest known substrate turnover rate of any P450, oxidizing arachidonic acid at a rate of 5000 min−1 [50]. The P450 and BMR domains retain their activities when they are separated and presented as two proteins and form regiospecific products as similar to those synthesized by the full-length enzyme, but the rate of product formation is lower [51]. BM3 is an important model for P450 research due to its self-sufficiency, high substrate turnover and solubility [46].

1.4.2 CYP116B2 (RhF)

Rhodococcus CYP116B2 (RhF) is a natural fusion enzyme comprising an N-terminal P450 domain fused to a C-terminal FMN-FeS didomain (FF) by a 22-amino-acid linker, as shown in Figure 4 [52]. Its physiological role is currently unclear because no natural substrate is known. It offers an ideal candidate P450 for direct protein evolution because it shows high substrate promiscuity, which is unusual for P450s. The enzyme catalyzes O-dealkylation, aromatic hydroxylation, olefin epoxidation and the asymmetric sulfoxidation of low-molecular-weight substituted aromatics [53].

Figure 4: Schematic representation of the natural P450-redox partner fusion enzyme RhF. FF, FMN-FeS didomain.
Figure 4:

Schematic representation of the natural P450-redox partner fusion enzyme RhF. FF, FMN-FeS didomain.

1.4.3 Artificial P450 fusion proteins

The engineering of P450s has focused on the optimization of the heme domain to improve or expand its catalytic performance, to overcome the need for a separate electron donor, to remove the need for NADPH, and to remove the membrane anchor [32]. Early strategies included enzymatic [54] and photochemical [55] cofactor regeneration, or chemical [56] and electrochemical [57] cofactor substitution. However, the construction of artificial fusion proteins to create self-sufficient P450s with integral reductases has received attention more recently [58]. Three major strategies have been developed: LICRED, PUPPET and “Molecular Lego”.

LICRED is a high-throughput method for the development of P450 fusion proteins in which a vector containing a ligation-independent cloning (LIC) site adjacent to the CYP116B2 reductase domain (RED) allows the rapid insertion of P450 genes [59], [60]. PUPPET is a platform in which three proliferating cell nuclear antigen (PCNA) fusion proteins are used to form a heterotrimer that recruits P450 into a complex with its electron donors, hence PUPPET refers to “proliferating cell nuclear antigen-utilized protein complex of P450 and its two electron transfer-related proteins” [61], [62]. Finally, the “Molecular Lego” method involves the combination of P450 enzymes with the BMR domain from the natural protein BM3 described above, to make recombinant enzymes with novel catalytic activities [63]. Soluble, self-sufficient human P450-BMR fusion enzymes have been created by removing the hydrophobic N-terminal membrane anchor domain from the insoluble human enzyme before fusion to the BMR sequence (Figure 5). This made it possible to purify the enzyme without detergents while retaining catalytic activities similar to wild-type P450s [64]. In addition to these three strategies, another method involves the exchange of shorter amino acid sequences between BM3 and novel insect P450s, an approach known as scanning chimeragenesis [65].

Figure 5: Schematic representation of the “Molecular Lego” principle for the construction of soluble, self-sufficient human P450-BMR fusion enzymes [32].
Figure 5:

Schematic representation of the “Molecular Lego” principle for the construction of soluble, self-sufficient human P450-BMR fusion enzymes [32].

2 Results

2.1 Construction of insect P450 fusion enzymes using the “Molecular Lego” approach

2.1.1 Cloning strategy

The “Molecular Lego” approach described above for human P450s is also ideal for the preparation of soluble and self-sufficient insect P450 fusion proteins because the BMR domain has a similar catalytic mechanism to the housefly CPR [1]. Two candidate insect P450s were therefore selected as shown in Table 2. In each case, codon usage was optimized for E. coli and the N-terminus was truncated by approximately 20 amino acids to remove the membrane anchor domain. It is important not to remove too many residues because this can compromise P450 activity, e.g. residues 21–82 of human CYP2E1 are necessary for heme incorporation and correct heme pocket folding [64]. Furthermore, the N-terminus should not include any hydrophobic amino acids, so the precise number of residues for removal was determined based on studies of human P450 enzymes used for the construction of fusion proteins [35]. This comparison suggested that residues 1–23 probably functioned as the membrane anchor and should be dispensable in terms of catalytic activity, so N-terminal trimming was carried out as shown in Figure 6. The linker determines the interaction between the P450 and BMR domains thus facilitating electron flow [69], [70], [71] and reducing its length by just six residues can abolish the activity of the fusion enzyme [49]. We therefore chose a linker containing 29 amino acids. The two P450-BMR fusion proteins were then prepared using the cloning strategy explained in Figure 7.

Table 2:

Characteristics of the insect P450s used for the preparation of fusion proteins.

P450OriginSubstratesGenBank
CYP6A1Musca domesticaHeptachlor, aldrin [66]M25367.1
CYP6G1Drosophila melanogasterImidacloprid, dichlorodiphenyltrichloroethane, methoxychlor [67], [68]AAF58557.1
Figure 6: N-terminal sequence alignment of the CYP6A1 and CYP6G1 proteins. The slashes highlighted in green indicate the start of the truncated sequences after removal of the membrane anchor domain. Black outline represents identity between residues. Gray outline represents similarity between residues.
Figure 6:

N-terminal sequence alignment of the CYP6A1 and CYP6G1 proteins. The slashes highlighted in green indicate the start of the truncated sequences after removal of the membrane anchor domain. Black outline represents identity between residues. Gray outline represents similarity between residues.

Figure 7: Cloning strategy used to construct the insect P450-BMR fusion enzymes. 6H, polyhistidine tag.
Figure 7:

Cloning strategy used to construct the insect P450-BMR fusion enzymes. 6H, polyhistidine tag.

2.1.2 Expression and purification strategy

The constructs were introduced into E. coli for the production of soluble recombinant protein, which was encouraged by inducing expression at 28°C. Buffers without detergents were used throughout purification, and all steps were carried out on ice or using cooled devices. The fusion proteins were captured by immobilized metal affinity chromatography (IMAC) and each eluted as a single peak in 250 mM imidazole buffer. The purity and integrity of the fusion proteins were then confirmed by sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE). The IMAC elution profile and SDS-PAGE analysis of the fusion proteins are shown in Figure 8.

Figure 8: Purification of CYP6G1-BMR and CYP6A1-BMR. (A) IMAC elution profile of CYP6G1-BMR and CYP6A1-BMR monitored at 280 nm and 410 nm. (B) SDS-PAGE analysis of purified CYP6G1-BMR. P, pellet after sonication and centrifugation; S, supernatant after sonication and centrifugation; 1–10, elution fractions. The dominant band migrating at 120 kDa in lanes 1–10 corresponds to the CYP6G1-BMR fusion protein.
Figure 8:

Purification of CYP6G1-BMR and CYP6A1-BMR. (A) IMAC elution profile of CYP6G1-BMR and CYP6A1-BMR monitored at 280 nm and 410 nm. (B) SDS-PAGE analysis of purified CYP6G1-BMR. P, pellet after sonication and centrifugation; S, supernatant after sonication and centrifugation; 1–10, elution fractions. The dominant band migrating at 120 kDa in lanes 1–10 corresponds to the CYP6G1-BMR fusion protein.

2.1.3 Activity of the fusion proteins

The activity of the reductase domain in the IMAC-purified fusion proteins was investigated by measuring NADPH turnover, which is an indirect indicator of catalytic activity and confirms correct protein folding. Reductase activity was measured in the presence of aldrin, the substrate for CYP6A1 [26], and imidacloprid, the substrate for CYP6G1 [68], and in controls without substrate. The assays were performed for 12 h with measurements taken every minute. Denatured proteins were used as negative controls. Figure 9 shows that the relative activity of purified CYP6A1-BMR reached 100%, as determined by NADPH consumption, in the presence and absence of substrate. The denatured CYP6A1-BMR negative control only achieved a residual relative activity of <20%. Similar results were achieved with CYP6G1-BMR, although the activity was lower due to the lower enzyme concentration. Thus, we were able to confirm BMR activity for both insect P450 fusion proteins.

Figure 9: BMR activity assay. NADPH turnover of CYP6A1-BMR (162 μg/ml) and CYP6G1-BMR (78 μg/ml) with aldrin (CYP6A1-BMR) or imidacloprid (CYP6G1-BMR) as substrates.
Figure 9:

BMR activity assay. NADPH turnover of CYP6A1-BMR (162 μg/ml) and CYP6G1-BMR (78 μg/ml) with aldrin (CYP6A1-BMR) or imidacloprid (CYP6G1-BMR) as substrates.

The activity of the P450 domain in the IMAC-purified fusion proteins was determined by measuring substrate turnover, which also provides evidence for the cooperation between the P450 and BMR domains. In the case of CYP6G1-BMR, we measured the conversion of the substrate imidacloprid to 4-hydroxyimidacloprid, 5-hydroxyimidacloprid, and 4,5-dihydroxyimidacloprid by ultra-high-performance liquid chromatography/mass spectrometry (UPLC-MS). In the case of CYP6A1-BMR, we measured the conversion of the substrate aldrin to dieldrin by gas chromatography-mass spectrometry (GC-MS). In both cases, we observed the loss of substrate during incubation, but the specific products were not detected.

2.1.4 Heme incorporation and substrate binding of the fusion proteins

To investigate the reason for the absence of P450 activity, we carried out further experiments to confirm heme incorporation into the P450 domain. The absorbance spectrum of CYP6G1-BMR showed a peak at 420 nm confirming the inclusion of a heme group, whereas the absorbance spectrum of CYP6A1-BMR did not show a peak at 420 nm suggesting the loss of heme after IMAC purification (Figure 10). Accordingly, CYP6A1-BMR was not able to catalyze the conversion of aldrin into dieldrin due to the absence of a functional catalytic center after purification. A photometric heme substrate-binding test was then carried out to determine whether the substrate bound to the catalytic heme domain of the purified fusion enzyme. There was no evidence of heme-substrate interaction, which may indicate poor access to the substrate-binding site resulting in the absence of a catalytically productive binding mode [72].

Figure 10: Absorbance spectra of purified BM3-CYP6A1 and BM3-CYP6G1 (300–500 nm). The peak at 420 nm indicates the presence of the heme group in the P450 domain of CYP6G1-BMR.
Figure 10:

Absorbance spectra of purified BM3-CYP6A1 and BM3-CYP6G1 (300–500 nm). The peak at 420 nm indicates the presence of the heme group in the P450 domain of CYP6G1-BMR.

3 Materials and methods

3.1 Gene synthesis

All genes were synthesized with a C-terminal His6-tag (GHHHHHH) and were codon optimized for expression in E. coli K12 (MWG Eurofins, Ebersberg, Germany). The genes encoding the full-length BM3 and insect P450s were flanked by BsaI sites for restriction digestion and ligation into the vector pASK-IBA33plus (IBA, Göttingen, Germany). XbaI and SalI sites were introduced into the BM3 gene to enable the removal of the endogenous P450 domain. The XbaI site was 41 bp upstream of the start codon, and the SalI site was inside the linker. Using this approach, three plasmids were generated: pASK-IBA33_BM3, pASK-IBA33_CYP6A1, and pASK-IBA33_CYP6G1.

3.2 Construction of the P450 fusion enzymes

The cloned insect P450 genes described above were amplified by PCR using the primers (MWG Eurofins, Ebersberg, Germany) listed in Table 3, adding XbaI and SalI sites at the 5′ and 3′ ends, respectively. The PCR products were double digested with XbaI and SalI (restriction enzymes from New England Biolabs, Frankfurt am Main, Germany) to create ~1500 bp P450 fragments, which were ligated into the 5113 bp vector backbone prepared by digesting pASK-IBA_BM3 with the same enzymes. This created the constructs pASK-IBA33_CYP6A1-BMR and pASK-IBA33_CYP6G1-BMR.

Table 3:

Oligonucleotide primers used for the preparation of insect P450 fusion constructs.

ConstructPrimer nameSequence
pASK-IBA33_CYP6A1-BMRCY6A1d20_fwd_XbaI5′-GGTATCTAGAATGAGCCGTTGGAACTTTGGGTATTGGAAAC

GTCGTGGTATTCCG-3′
CYP6A1_rev_linker+SalI5′-CCGGTCGACGGACTAGGGATCCCTCCGAGGGGAATTTTCT

TGCTTTTGATTTTCTTGCGAATTGC-3′
pASK-IBA33_CYP6G1-BMRCYP6G1d20_fwd_XbaI5′-GGTATCTAGAATGAGCCGCAATCACTCATACTGGCAGCGTA

AAGGGATTCCGTACATTCCG-3′
CYP6G1_rev_linker+SalI5′-CCGGTCGACGGACTAGGGATCCCTCCGAGGGGAATTTTCT

TGCTTTGAAGCGACGGAGCTGATTG-3′

3.3 Expression

Terrific broth medium containing 100 μg/ml carbenicillin (Carl Roth, Karlsruhe, Germany), 0.5 mM 5-aminolevulinic acid (Sigma-Aldrich, Schnelldorf, Germany), 0.5 mM thiamine (Carl Roth, Karlsruhe, Germany), 0.25 μg/ml FeCl3 (Sigma-Aldrich, Schnelldorf, Germany), 1 μM riboflavin (Sigma-Aldrich, Schnelldorf, Germany), and 160 mM d(+)glucose (Carl Roth, Karlsruhe, Germany) was inoculated with a fresh 16 h overnight lysogeny broth culture (Carl Roth, Karlsruhe, Germany) of E. coli (Invitrogen, Carlsbad, USA) carrying the vectors described above. The expression of genes carried on plasmid pASK-IBA33plus was induced with 200 μg/l anhydrotetracycline (IBA, Göttingen, Germany). Protein production was triggered by anhydrotetracycline once the OD600 reached 0.5. Cells were grown at 28°C for 4 h and harvested by centrifugation before lysis.

3.4 IMAC

The cell pellet was resuspended in 40–60 ml lysis buffer [pH 7.5, 30 mM Tris-HCl (Carl Roth, Karlsruhe, Germany), 100 mM NaCl (Carl Roth, Karlsruhe, Germany), 20% glycerol (v/v) (Carl Roth, Karlsruhe, Germany)], and the lysed cells were centrifuged at 75,600×g for 1 h at 10°C. The supernatant was loaded on a freshly packed and conditioned 3 ml Ni-NTA (Macherey-Nagel, Weilmünster, Germany) column, connected to an SE04 system (ECOM, Prague, Czech Republic) maintained at 4°C, at a flow rate of 1 ml/min. The column was washed with buffer A [30 mM Tris-HCl, 100 mM NaCl, 20% glycerol (v/v), 30 mM imidazole (Sigma-Aldrich, Schnelldorf, Germany), pH 7.5] at 3 ml/min until the baseline was reached. The enzyme was eluted with 250 mM imidazole at 3 ml/min.

3.5 Activity assay

Each reaction mixture contained 78 μg/ml (CYP6G1-BMR) or 162 μg/ml (CYP6A1-BMR) purified fusion enzyme, 0.5 mM NADPH (Sigma-Aldrich, Schnelldorf, Germany), 0.1 mM substrate, 25 mM Tris-HCl (pH 7.5), and 20% glycerol (v/v) in a total volume of 200 μl. As controls, samples were prepared without NADPH, without substrate and with heat-denatured enzyme. The oxidation of NADPH was monitored at 340 nm on an Eon microplate reader (Biotek, Bad Friedrichshall, Germany) with one measurement every minute for 12 h.

3.6 UPLC-MS

Each reaction mixture comprised 78 μg/ml purified CYP6G1-BMR, 100 μg/ml imidacloprid (Sigma-Aldrich, Schnelldorf, Germany), 0.5 mM NADPH and 25 mM Tris-HCl (pH 7.5). After incubation at room temperature for 25 h the samples were mixed 1:1 with acetonitrile and transported on ice to Chromicent (Berlin, Germany)for UPLC-MS analysis according to [73]. A Waters (Eschborn, Germany) UPLC instrument was equipped with an AQUITY HSS T3 column (100×2.1 mm internal diameter, 1.8 μm particle size) coupled to a Waters Xevo TQ-S micro mass spectrometer. Water with 0.1% formic acid (Sigma-Aldrich, Schnelldorf, Germany) was used as solvent A and acetonitrile (Sigma-Aldrich, Schnelldorf, Germany) as solvent B. The injection volume was 5 μl. Each separation took 17 min at a flow rate of 0.5 ml/min, starting at 95% solvent A (held for 3 min) and decreasing to 40% solvent A in 12 min (held for 2 min). MS was operated in positive electrospray ionization mode. The source temperature was set at 120°C with nitrogen flow rates of 20 l/h for the cone gas and 1000 l/h for the desolvation gas. The desolvation temperature was 600°C. Standard curves of imidacloprid and imidacloprid-olefin (Dr. Ehrenstorfer, Augsburg, Germany) (3.125 μg/ml, 6.25 μg/ml, 12.5 μg/ml, 25 μg/ml, 50 μg/ml and 100 μg/ml) were prepared in triplicate before the samples were processed. Imidacloprid and imidacloprid-olefin were quantified using the corresponding standard curve.

3.7 Photometric heme substrate-binding test

Samples for the photometric heme substrate-binding test were transferred to 1 ml cuvettes for analysis in a photometer (Specord 210, Analytik Jena AG, Jena, Germany). The baseline was determined using 400 μl of the appropriate buffer plus 40 μl dimethylsulfoxide (DMSO) (Carl Roth, Karlsruhe, Germany) as the substrate solvent. The first measurement from 300 to 500 nm was performed with 400 μl of an undiluted elution sample of CYP6G1-BMR, or denatured CYP6G1-BMR and 40 μl DMSO. After this measurement, 400 μl of the same samples were mixed with 40 μl 1 mg/ml imidacloprid in DMSO and the measurement was repeated. The results were compared to see if the peak shifted from 425 nm to another wavelength.

4 Conclusions and outlook

The enormous technological potential of insect P450 can only be realised if sufficient quantities of high-quality active enzymes are made available for molecular and biochemical characterisation. We applied the “Molecular Lego” strategy by constructing the insect P450-BMR fusion enzymes for expression as soluble proteins in E. coli, allowing their purification without detergents, but in vitro P450 activity could not be detected even though the BMR domain was functional in both candidates.

Several strategies remain available to improve the activity of our insect P450-BMR fusion proteins. The fusion proteins could be expressed in other heterologous systems such as insect cells or yeast, which may provide a more appropriate molecular environment. Furthermore, we used vector pASK-IBA33plus because it has achieved the soluble expression of other insect enzymes in E. coli [74], but switching to the pCW ori+ system may generate better results because the latter is the best characterized vector for P450 expression [75]. The linker region could be modified by changing its length and/or composition to optimize the interaction between the BMR and P450 domains. The flexibility of the linker and the electron transfer rate between the BMR and P450 domains could also be improved by random or site-directed mutagenesis. In addition, the elimination of potential protease cleavage sites could enhance the stability of the fusion protein. Increasing the length of the N-terminus may improve enzyme activity because excessive truncation can inhibit heme incorporation and correct heme pocket folding [64]. The versatile RhF reductase domain may also prove useful because this has been previously successfully combined with P450s from other sources [76], [77]. X-ray diffraction data provide a more rational basis for the design of P450-CPR fusion proteins, but no crystal structures are yet available for the insect P450s [64]. Our results provide a solid foundation for future work on insect P450s, including their exploitation for industrial processes.

  1. Funding: Hessen State Ministry of Higher Education, Research and the Arts (HMWK), (Grant/Award Number: LOEWE-Center for Insect Biotechnology and Bioresou).

References

1. Feyereisen R. Insect P450 enzymes. Annu Rev Entomol 1999;44:507–33.10.1146/annurev.ento.44.1.507Search in Google Scholar

2. Guengerich FP. In: Ortiz de Montellano P, editor. Human cytochrome P450 enzymes. Cytochrome P450. Boston, MA: Springer, 2005:377–530.Search in Google Scholar

3. Hille UE, Hu Q, Pinto-Bazurco Mendieta MAE, Bartels M, Vock CA, Lauterbach T, et al. Steroidogenic cytochrome P450 (CYP) enzymes as drug targets: combining substructures of known CYP inhibitors leads to compounds with different inhibitory profile. Comptes Rendus Chimie 2009;12:1117–26.10.1016/j.crci.2009.03.007Search in Google Scholar

4. Kelly S, Kelly D, Jackson C, Warrilow AS, Lamb D. The diversity and importance of microbial cytochromes P450. In: Ortiz de Montellano P, editor. Cytochrome P450. Boston, MA: Springer, 2005:585–617.10.1007/0-387-27447-2_13Search in Google Scholar

5. Klingenberg M. Pigments of rat liver microsomes. Arch Biochem Biophys 1958;75:376–86.10.1016/0003-9861(58)90436-3Search in Google Scholar

6. Omura T, Sato R. The carbon monoxide-binding pigment of liver microsomes: I. evidence for its hemoprotein nature. J Biol Chem 1964;239:2370–8.10.1016/S0021-9258(20)82244-3Search in Google Scholar

7. Omura T, Sato R. The carbon monoxide-binding pigment of liver microsomes: II. solubilization, purification, and properties. J Biol Chem 1964;239:2379–85.10.1016/S0021-9258(20)82245-5Search in Google Scholar

8. Nelson DR. A world of cytochrome P450s. Philos Trans R Soc Lond B Biol Sci 2013;368:19.10.1098/rstb.2012.0430Search in Google Scholar

9. Nebert DW, Adesnik M, Coon MJ, Estabrook RW, Gonzalez FJ, Guengerich FP, et al. The P450 gene superfamily: recommended nomenclature. DNA 1987;6:1–11.10.1089/dna.1987.6.1Search in Google Scholar

10. Hedegaard J, Gunsalus IC. Mixed function oxidation: IV. An induced methylene hydroxylase in camphor oxidation. J Biol Chem 1965;240:4038–43.10.1016/S0021-9258(18)97147-4Search in Google Scholar

11. Dus K, Katagiri M, Yu CA, Erbes DL, Gunsalus IC. Chemical characterization of cytochrome P-450cam. Biochem Biophys Res Commun 1970;40:1423–30.10.1016/0006-291X(70)90026-4Search in Google Scholar

12. Poulos T, Johnson E. Structures of cytochrome P450 enzymes. In: Ortiz de Montellano P, editor. Cytochrome P450. Boston, MA: Springer, 2005:87–114.10.1007/0-387-27447-2_3Search in Google Scholar

13. Guengerich FP. Mechanisms of cytochrome P450 substrate oxidation: minireview. J Biochem Mol Toxicol 2007;21:163–8.10.1002/jbt.20174Search in Google Scholar PubMed

14. Ortiz de Montellano PR. Hydrocarbon hydroxylation by cytochrome P450 enzymes. Chem Rev 2010;110:932–48.10.1021/cr9002193Search in Google Scholar PubMed PubMed Central

15. Fasan R. Tuning P450 enzymes as oxidation catalysts. ACS Catal 2012;2:647–66.10.1021/cs300001xSearch in Google Scholar

16. Lamb DC, Waterman MR. Unusual properties of the cytochrome P450 superfamily. Philos Trans R Soc Lond B Biol Sci 2013;368.10.1098/rstb.2012.0434Search in Google Scholar PubMed PubMed Central

17. Werck-Reichhart D, Feyereisen R. Cytochromes P450: a success story. Genome Biol 2000;1:3003.1–3003.9.10.1186/gb-2000-1-6-reviews3003Search in Google Scholar PubMed PubMed Central

18. Guengerich FP. Common and uncommon cytochrome P450 reactions related to metabolism and chemical toxicity. Chem Res Toxicol 2001;14:611–50.10.1021/tx0002583Search in Google Scholar PubMed

19. Estabrook RW. The remarkable P450s: a historical overview of these versatile hemeprotein catalysts. FASEB J 1996;10:202–4.10.1096/fasebj.10.2.8641552Search in Google Scholar PubMed

20. Scott JG, Wen Z. Cytochromes P450 of insects: the tip of the iceberg. Pest Manag Sci 2001;57:958–67.10.1002/ps.354Search in Google Scholar PubMed

21. Rewitz KF, Rybczynski R, Warren JT, Gilbert LI. The Halloween genes code for cytochrome P450 enzymes mediating synthesis of the insect moulting hormone. Biochem Soc Trans 2006;34(Pt 6):1256–60.10.1042/BST0341256Search in Google Scholar PubMed

22. Scott JG. Cytochromes P450 and insecticide resistance. Insect Biochem Mol Biol 1999;29:757–77.10.1016/S0965-1748(99)00038-7Search in Google Scholar

23. Scott JG, Liu N, Wen Z. Insect cytochromes P450: diversity, insecticide resistance and tolerance to plant toxins. Comp Biochem Physiol Part C: Pharmacol Toxicol Endocrinol 1998;121: 147–55.10.1016/S0742-8413(98)10035-XSearch in Google Scholar

24. Feyereisen R. Insect P450 inhibitors and insecticides: challenges and opportunities. Pest Manag Sci 2015;71: 793–800.10.1002/ps.3895Search in Google Scholar

25. Wilson TG, Hodgson E. Microsomal NADPH*-cytochrome c reductase from the housefly, Musca domestica solubilization and purification. Insect Biochem 1971;1:19–26.10.1016/0020-1790(71)90018-7Search in Google Scholar

26. Andersen JF, Utermohlen JG, Feyereisen R. Expression of housefly CYP6A1 and NADPH-cytochrome P450 reductase in Escherichia coli and reconstitution of an insecticide-metabolizing P450 system. Biochemistry 1994;33:2171–7.10.1021/bi00174a025Search in Google Scholar PubMed

27. Andersen JF, Walding JK, Evans PH, Bowers WS, Feyereisen R. Substrate specificity for the epoxidation of terpenoids and active site topology of house fly cytochrome P450 6A1. Chem Res Toxicol 1997;10:156–64.10.1021/tx9601162Search in Google Scholar PubMed

28. Cheesman MJ, Traylor MJ, Hilton ME, Richards KE, Taylor MC, Daborn PJ, et al. Soluble and membrane-bound Drosophila melanogaster CYP6G1 expressed in Escherichia coli: purification, activity, and binding properties toward multiple pesticides. Insect Biochem Mol Biol 2013;43:455–65.10.1016/j.ibmb.2013.02.003Search in Google Scholar PubMed

29. Howard RW, Blomquist GJ. Ecological, behavioral, and biochemical aspects of insect hydrocarbons. Annu Rev Entomol 2005;50:371–93.10.1146/annurev.ento.50.071803.130359Search in Google Scholar PubMed

30. Qiu Y, Tittiger C, Wicker-Thomas C, Le Goff G, Young S, Wajnberg E, et al. An insect-specific P450 oxidative decarbonylase for cuticular hydrocarbon biosynthesis. Proc Natl Acad Sci USA 2012;109:14858–63.10.1073/pnas.1208650109Search in Google Scholar PubMed PubMed Central

31. O’Reilly E, Kohler V, Flitsch SL, Turner NJ. Cytochromes P450 as useful biocatalysts: addressing the limitations. Chem Commun 2011;47:2490–501.10.1039/c0cc03165hSearch in Google Scholar PubMed

32. Sadeghi SJ, Gilardi G. Chimeric P450 enzymes: activity of artificial redox fusions driven by different reductases for biotechnological applications. Biotechnol Appl Biochem 2013;60:102–10.10.1002/bab.1086Search in Google Scholar

33. Klenk JM, Nebel BA, Porter JL, Kulig JK, Hussain SA, Richter SM, et al. The self-sufficient P450 RhF expressed in a whole cell system selectively catalyses the 5-hydroxylation of diclofenac. Biotechnol J 2017;12:1600520.10.1002/biot.201600520Search in Google Scholar

34. Kempf AC, Zanger UM, Meyer UA. Truncated Human P450 2D6P: expression in Escherichia coli, Ni2+-chelate affinity purification, and characterization of solubility and aggregation. Arch Biochem Biophys 1995;321:277–88.10.1006/abbi.1995.1396Search in Google Scholar

35. Dodhia V, Fantuzzi A, Gilardi G. Engineering human cytochrome P450 enzymes into catalytically self-sufficient chimeras using molecular Lego. J Biol Inorg Chem 2006;11:903–16.10.1007/s00775-006-0144-3Search in Google Scholar

36. Bernhardt R, Urlacher VB. Cytochromes P450 as promising catalysts for biotechnological application: chances and limitations. Appl Microbiol Biotechnol 2014;98:6185–203.10.1007/s00253-014-5767-7Search in Google Scholar

37. Lundemo MT, Woodley JM. Guidelines for development and implementation of biocatalytic P450 processes. Appl Microbiol Biotechnol 2015;99:2465–83.10.1007/s00253-015-6403-xSearch in Google Scholar

38. Zhang L, Liu X, Wang C, Liu X, Cheng G, Wu Y. Expression, purification and direct eletrochemistry of cytochrome P450 6A1 from the house fly, Musca domestica. Protein Expr Purif 2010;71: 74–8.10.1016/j.pep.2009.12.008Search in Google Scholar

39. Nakahara K, Shoun H. N-terminal processing and amino acid sequence of two isoforms of nitric oxide reductase cytochrome P450nor from Fusarium oxysporum. J Biochem 1996;120: 1082–7.10.1093/oxfordjournals.jbchem.a021525Search in Google Scholar

40. Narhi LO, Fulco AJ. Characterization of a catalytically self-sufficient 119,000-dalton cytochrome P-450 monooxygenase induced by barbiturates in Bacillus megaterium. J Biol Chem 1986;261:7160–9.10.1016/S0021-9258(17)38369-2Search in Google Scholar

41. Roberts GA, Grogan G, Greter A, Flitsch SL, Turner NJ. Identification of a new class of cytochrome P450 from a Rhodococcus sp. J Bacteriol 2002;184:3898–908.10.1128/JB.184.14.3898-3908.2002Search in Google Scholar PubMed PubMed Central

42. Jackson CJ, Lamb DC, Marczylo TH, Warrilow AGS, Manning NJ, Lowe DJ, et al. A novel sterol 14α-demethylase/ferredoxin fusion protein (MCCYP51FX) from Methylococcus capsulatus represents a new class of the cytochrome P450 superfamily. J Biol Chem 2002;277:46959–65.10.1074/jbc.M203523200Search in Google Scholar

43. Rylott EL, Jackson RG, Sabbadin F, Seth-Smith HM, Edwards J, Chong CS, et al. The explosive-degrading cytochrome P450 XplA: biochemistry, structural features and prospects for bioremediation. Biochim Biophys Acta 2011;1814:230–6.10.1016/j.bbapap.2010.07.004Search in Google Scholar

44. Munro AW, Girvan HM, McLean KJ. Cytochrome P450–redox partner fusion enzymes. Biochim Biophys Acta 2007;1770:345–59.10.1016/j.bbagen.2006.08.018Search in Google Scholar

45. Lamb DC, Lei L, Warrilow AGS, Lepesheva GI, Mullins JGL, Waterman MR, et al. The first virally encoded cytochrome P450. J Virol 2009;83:8266–9.10.1128/JVI.00289-09Search in Google Scholar

46. Munro AW, Leys DG, McLean KJ, Marshall KR, Ost TW, Daff S, et al. P450 BM3: the very model of a modern flavocytochrome. Trends Biochem Sci 2002;27:250–7.10.1016/S0968-0004(02)02086-8Search in Google Scholar

47. Ravichandran K, Boddupalli S, Hasermann C, Peterson J, Deisenhofer J. Crystal structure of hemoprotein domain of P450BM-3, a prototype for microsomal P450’s. Science 1993;261:731–6.10.1126/science.8342039Search in Google Scholar

48. Li HY, Darwish K, Poulos TL. Characterization of recombinant Bacillus megaterium cytochrome P-450 BM-3 and its two functional domains. J Biol Chem 1991;266:11909–14.10.1016/S0021-9258(18)99044-7Search in Google Scholar

49. Govindaraj S, Poulos TL. Role of the linker region connecting the reductase and heme domains in cytochrome P450BM-3. Biochemistry 1995;34:11221–6.10.1021/bi00035a031Search in Google Scholar PubMed

50. Noble MA, Miles CS, Chapman SK, Lysek DA, MacKay AC, Reid GA, et al. Roles of key active-site residues in flavocytochrome P450 BM3. Biochem J 1999;339(Pt 2):371–9.10.1042/bj3390371Search in Google Scholar

51. Govindaraj S, Poulos TL. The domain architecture of cytochrome P450BM-3. J Biol Chem 1997;272:7915–21.10.1074/jbc.272.12.7915Search in Google Scholar PubMed

52. Hunter DJB, Roberts GA, Ost TWB, White JH, Müller S, Turner NJ, et al. Analysis of the domain properties of the novel cytochrome P450 RhF. FEBS Lett 2005;579:2215–20.10.1016/j.febslet.2005.03.016Search in Google Scholar

53. O’Reilly E, Corbett M, Hussain S, Kelly PP, Richardson D, Flitsch SL, et al. Substrate promiscuity of cytochrome P450 RhF. Catal Sci Tech 2013;3:1490–2.10.1039/c3cy00091eSearch in Google Scholar

54. Falck JR, Reddy YK, Haines DC, Reddy KM, Krishna UM, Graham S, et al. Practical, enantiospecific syntheses of 14,15-EET and leukotoxin B (vernolic acid). Tetrahedron Lett 2001;42:4131–3.10.1016/S0040-4039(01)00694-3Search in Google Scholar

55. Shumyantseva VV, Bulko TV, Schmid RD, Archakov AI. Photochemical properties of a riboflavins/cytochrome P450 2B4 complex. Biosens Bioelectron 2002;17:233–8.10.1016/S0956-5663(01)00181-6Search in Google Scholar

56. Shumyantseva VV, Bulko TV, Archakov AI. Electrochemical reduction of cytochrome P450 as an approach to the construction of biosensors and bioreactors. J Inorg Biochem 2005;99:1051–63.10.1016/j.jinorgbio.2005.01.014Search in Google Scholar PubMed

57. Fang X, Halpert JR. Dithionite-supported hydroxylation of palmitic acid by cytochrome P450BM-3. Drug Metab Dispos 1996;24:1282–5.Search in Google Scholar

58. McLean KJ, Girvan HM, Munro AW. Cytochrome P450/redox partner fusion enzymes: biotechnological and toxicological prospects. Expert Opin Drug Metab Toxicol 2007;3:847–63.10.1517/17425255.3.6.847Search in Google Scholar PubMed

59. Sabbadin F, Hyde R, Robin A, Hilgarth EM, Delenne M, Flitsch S, et al. LICRED: a versatile drop-in vector for rapid generation of redox-self-sufficient cytochrome P450s. Chembiochem 2010;11:987–94.10.1007/978-1-62703-321-3_20Search in Google Scholar PubMed

60. Sabbadin F, Grogan G, Bruce NC. LICRED: a versatile drop-in vector for rapid generation of redox-self-sufficient cytochromes P450. In: Phillips RI, Shephard AE, Ortiz de Montellano RP, editors. Cytochrome P450 Protocols. Totowa, NJ: Humana Press, 2013:239–49.10.1007/978-1-62703-321-3_20Search in Google Scholar

61. Haga T, Hirakawa H, Nagamune T. Fine tuning of spatial arrangement of enzymes in a PCNA-mediated multienzyme complex using a rigid poly-L-proline linker. PLoS One 2013;8:e75114.10.1371/journal.pone.0075114Search in Google Scholar PubMed PubMed Central

62. Hirakawa H, Nagamune T. Molecular assembly of P450 with ferredoxin and ferredoxin reductase by fusion to PCNA. ChemBioChem 2010;11:1517–20.10.1002/cbic.201000226Search in Google Scholar PubMed

63. Gilardi G, Meharenna YT, Tsotsou GE, Sadeghi SJ, Fairhead M, Giannini S. Molecular Lego: design of molecular assemblies of P450 enzymes for nanobiotechnology. Biosens Bioelectron 2002;17:133–45.10.1016/S0956-5663(01)00286-XSearch in Google Scholar

64. Fairhead M, Giannini S, Gillam EJ, Gilardi G. Functional characterisation of an engineered multidomain human P450 2E1 by molecular Lego. J Biol Inorg Chem 2005;10:842–53.10.1007/s00775-005-0033-1Search in Google Scholar PubMed

65. Chen CK, Berry RE, Shokhireva T, Murataliev MB, Zhang H, Walker FA. Scanning chimeragenesis: the approach used to change the substrate selectivity of fatty acid monooxygenase CYP102A1 to that of terpene omega-hydroxylase CYP4C7. J Biol Inorg Chem 2010;15:159–74.10.1007/s00775-009-0580-ySearch in Google Scholar PubMed

66. Feyereisen R, Andersen JF, Cariño FA, Cohen MB, Koener JF. Cytochrome P450 in the house fly: structure, catalytic activity and regulation of expression of CYP6 A1 in an insecticide-resistant strain. Pest Manag Sci 1995;43:233–9.10.1002/ps.2780430309Search in Google Scholar

67. Joußen N, Heckel DG, Haas M, Schuphan I, Schmidt B. Metabolism of imidacloprid and DDT by P450 CYP6G1 expressed in cell cultures of Nicotiana tabacum suggests detoxification of these insecticides in Cyp6g1-overexpressing strains of Drosophila melanogaster, leading to resistance. Pest Manag Sci 2008;64:65–73.10.1002/ps.1472Search in Google Scholar PubMed

68. Joußen N, Schuphan I, Schmidt B. Metabolism of methoxychlor by the P450-monooxygenase CYP6G1 involved in insecticide resistance of Drosophila melanogaster after expression in cell cultures of Nicotiana tabacum. Chem Biodivers 2010;7:722–35.10.1002/cbdv.200900020Search in Google Scholar PubMed

69. Govindaraj S, Poulos TL. Probing the structure of the linker connecting the reductase and heme domains of cytochrome P450BM-3 using site-directed mutagenesis. Protein Sci 1996;5:1389–93.10.1002/pro.5560050717Search in Google Scholar PubMed PubMed Central

70. Chen X, Zaro JL, Shen WC. Fusion protein linkers: property, design and functionality. Adv Drug Deliv Rev 2013;65:1357–69.10.1016/j.addr.2012.09.039Search in Google Scholar PubMed PubMed Central

71. Hoffmann SM, Weissenborn MJ, Gricman Ł, Notonier S, Pleiss J, Hauer B. The impact of linker length on P450 fusion constructs: activity, stability and coupling. ChemCatChem 2016;8:1591–7.10.1002/cctc.201501397Search in Google Scholar

72. Zhang W, Liu Y, Yan J, Cao S, Bai F, Yang Y, et al. New reactions and products resulting from alternative interactions between the P450 enzyme and redox partners. J Am Chem Soc 2014;136:3640–6.10.1021/ja4130302Search in Google Scholar PubMed PubMed Central

73. Kamel A. Refined methodology for the determination of neonicotinoid pesticides and their metabolites in honey bees and bee products by liquid chromatography–tandem mass spectrometry (LC-MS/MS). J Agric and Food Chem 2010;58:5926–31.10.1021/jf904120nSearch in Google Scholar PubMed

74. Pöppel A-K, Kahl M, Baumann A, Wiesner J, Gökçen A, Beckert A, et al. A Jonah-like chymotrypsin from the therapeutic maggot Lucilia sericata plays a role in wound debridement and coagulation. Insect Biochem Mol Biol 2016;70:138–47.10.1016/j.ibmb.2015.11.012Search in Google Scholar PubMed

75. Pritchard MP, McLaughlin L, Friedberg T. Establishment of functional human cytochrome P450 monooxygenase systems in Escherichia coli. In: Phillips IR, Shephard EA, editors. Cytochrome P450 Protocols. Totowa, NJ: Humana Press, 2006:19–29.10.1385/1-59259-998-2:19Search in Google Scholar

76. Robin A, Köhler V, Jones A, Ali A, Kelly PP, O’Reilly E, et al. Chimeric self-sufficient P450cam-RhFRed biocatalysts with broad substrate scope. Beilstein J Org Chem 2011;7:1494–8.10.3762/bjoc.7.173Search in Google Scholar PubMed PubMed Central

77. Nodate M, Kubota M, Misawa N. Functional expression system for cytochrome P450 genes using the reductase domain of self-sufficient P450RhF from Rhodococcus sp. NCIMB 9784. Appl Microbiol Biotechnol 2006;71:455–62.10.1007/s00253-005-0147-ySearch in Google Scholar PubMed

Received: 2017-3-8
Revised: 2017-5-26
Accepted: 2017-6-8
Published Online: 2017-9-4
Published in Print: 2017-9-26

©2017 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 29.3.2024 from https://www.degruyter.com/document/doi/10.1515/znc-2017-0041/html
Scroll to top button