Skip to main content
Applied and Environmental Microbiology logoLink to Applied and Environmental Microbiology
. 2018 Jul 2;84(14):e02777-17. doi: 10.1128/AEM.02777-17

Engineered 3-Ketosteroid 9α-Hydroxylases in Mycobacterium neoaurum: an Efficient Platform for Production of Steroid Drugs

Hao-Hao Liu a, Li-Qin Xu a, Kang Yao a, Liang-Bin Xiong a, Xin-Yi Tao a, Min Liu a, Feng-Qing Wang a,, Dong-Zhi Wei a,
Editor: Isaac Cannb
PMCID: PMC6029100  PMID: 29728384

Steroidal drugs are widely used for anti-inflammation, anti-tumor action, endocrine regulation, and fertility management, among other uses. The two main starting materials for the industrial synthesis of steroid drugs are phytosterol and diosgenin. The phytosterol processing is carried out by microbial transformation, which is thought to be superior to the diosgenin processing by chemical conversions, given its simple and environmentally friendly process. However, diosgenin has long been used as the primary starting material instead of phytosterol. This is in response to challenges in developing efficient microbial strains for industrial phytosterol transformation, which stem from complex metabolic processes that feature many currently unclear details. In this study, we identified two oxygenase homologues of 3-ketosteroid-9α-hydroxylase, KshA1N and KshA2N, in M. neoaurum and demonstrated their crucial role in determining the yield and variety of products from phytosterol transformation. This work has practical value in developing industrial strains for phytosterol biotransformation.

KEYWORDS: 3-ketosteroid-9α-hydroxylase; androst-1,4-diene-3,17-dione (ADD); 9α-hydroxy-4-androstene-3,17-dione (9-OHAD); mycobacterium; Mn25795; sterols; 9α-hydroxy-4-androstene-3,17-dione; androst-1,4-diene-3,17-dione; sterol

ABSTRACT

3-Ketosteroid 9α-hydroxylase (Ksh) consists of a terminal oxygenase (KshA) and a ferredoxin reductase and is indispensable in the cleavage of steroid nucleus in microorganisms. The activities of Kshs are crucial factors in determining the yield and distribution of products in the biotechnological transformation of sterols in industrial applications. In this study, two KshA homologues, KshA1N and KshA2N, were characterized and further engineered in a sterol-digesting strain, Mycobacterium neoaurum ATCC 25795, to construct androstenone-producing strains. kshA1N is a member of the gene cluster encoding sterol catabolism enzymes, and its transcription exhibited a 4.7-fold increase under cholesterol induction. Furthermore, null mutation of kshA1N led to the stable accumulation of androst-4-ene-3,17-dione (AD) and androst-1,4-diene-3,17-dione (ADD). We determined kshA2N to be a redundant form of kshA1N. Through a combined modification of kshA1N, kshA2N, and other key genes involved in the metabolism of sterols, we constructed a high-yield ADD-producing strain that could produce 9.36 g liter−1 ADD from the transformation of 20 g liter−1 phytosterols in 168 h. Moreover, we improved a previously established 9α-hydroxy-AD-producing strain via the overexpression of a mutant KshA1N that had enhanced Ksh activity. Genetic engineering allowed the new strain to produce 11.7 g liter−1 9α-hydroxy-4-androstene-3,17-dione (9-OHAD) from the transformation of 20.0 g liter−1 phytosterol in 120 h.

IMPORTANCE Steroidal drugs are widely used for anti-inflammation, anti-tumor action, endocrine regulation, and fertility management, among other uses. The two main starting materials for the industrial synthesis of steroid drugs are phytosterol and diosgenin. The phytosterol processing is carried out by microbial transformation, which is thought to be superior to the diosgenin processing by chemical conversions, given its simple and environmentally friendly process. However, diosgenin has long been used as the primary starting material instead of phytosterol. This is in response to challenges in developing efficient microbial strains for industrial phytosterol transformation, which stem from complex metabolic processes that feature many currently unclear details. In this study, we identified two oxygenase homologues of 3-ketosteroid-9α-hydroxylase, KshA1N and KshA2N, in M. neoaurum and demonstrated their crucial role in determining the yield and variety of products from phytosterol transformation. This work has practical value in developing industrial strains for phytosterol biotransformation.

INTRODUCTION

Some microorganisms can naturally use sterols as carbon and energy sources. During the microbial metabolism of sterols, several metabolic intermediates can be readily used to synthesize useful steroid drugs, including androst-4-ene-3,17-dione (AD), androst-1,4-diene-3,17-dione (ADD), 9α-hydroxy AD (9α-OHAD), and 22-hydroxy-23,24-bisnorchol-4-en-3-one (4-HBC), among others. Thus, an important means for the commercial production of steroid drugs has been developed based on the biotechnological modification of these microbial metabolic processes, primarily in the mycobacteria (1, 2). To develop a strain that can produce one of these valuable metabolites in high yield, several enzymatic steps must be considered. The 9α-hydroxylation and Δ1-dehydrogenation of 3-ketosteroids, which are carried out by 3-ketosteroid 9α-hydroxylases (Kshs) and 3-ketosteroid Δ1-dehydrogenases (KstDs), respectively, are two key steps that must be precisely controlled because of their essential roles in the selective formation and further degradation of target metabolites (3, 4). According to the proposed pathway of sterol catabolism in these microorganisms (Fig. 1a), genetic inactivation of Kshs or KstDs will block the degradation of steroid nucleus and thereby lead to the accumulation of steroidal intermediates, with three possible nucleus structures: AD, ADD, or 9α-OHAD (Fig. 1b). Theoretically, full inactivation of Kshs can achieve the production of ADD while maintaining sufficient activation of other enzymes involved in sterol metabolism, especially KstDs. Likewise, complete inactivation of KstDs can result in the production of 9α-OHAD while maintaining sufficient activation of other enzymes, especially Kshs. The inactivation of both Kshs and KstDs can result in AD as the main product. Similarly, various C22 metabolites also can be achieved by combined modification of Kshs, KstDs, and other key enzymes involved in side chain degradation (Fig. 1b). Therefore, Kshs and KstDs are important targets for modification by random mutagenesis or genetic engineering to develop microorganisms that can transform sterols to valuable steroidal intermediates.

FIG 1.

FIG 1

Role of Ksh in the sterol catabolism pathway under aerobic conditions. (a) Metabolism of the steroid nucleus. (b) Proposed catabolic pathway of sterols (e.g., β-sitosterol). C19 metabolites are shown in the blue box, and C22 metabolites are in the green box. The conversion from 22HOBNC-CoA to AD was designated the AD pathway (blue arrows), and the conversion from 22HOBNC-CoA to 4-HBC was designated the HBC pathway (green arrows). Abbreviations: 22OBNC-CoA, 3,22-dioxo-25,26-bisnorchol-4-ene-24-oyl CoA; 22HOBNC-CoA, 22-hydroxy-3-oxo-25,26-bisnorchol-4-en-24-oyl CoA; 4-BNC, 3-oxo-23,24-bisnorchol-4-en-22-oic acid; 4-BNC-CoA, 3-oxo-23,24-bisnorchol-4-en-22-oyl-coenzyme A thioester; AD, 4-androstene-3,17-dione; ADD, 1,4-androstadiene-3,17-dione; T, testosterone; DHT, boldenone; 9-OHAD, 9α-hydroxy-4-androstene-3,17-dione; 9-OHADD, 9α-hydroxy-1,4-androstadiene-3,17-dione; 4-HBC, 22-hydroxy-23,24-bisnorchol-ene-3-one; 1,4-HBC, 22-hydroxy-23,24-bisnorchol-1,4-dien-3-one; 9-OHHBC, 9,22-dihydroxy-23,24-bisnorchol-4-ene-3-one.

Ksh is a two-component oxygenase, including a Rieske oxygenase (KshA) and a ferredoxin reductase (KshB) that consumes O2 and 3-ketosteroids as substrates and NAD as the electron donor (3, 5). The first Ksh system was identified and molecularly characterized in Rhodococcus erythropolis SQ1, which confirmed the previously speculated catabolic processing of steroid nucleus and revealed the physiological importance of the Ksh system (6). The second Ksh system was subsequently isolated and identified in this strain (7). The cholesterol metabolism in Mycobacterium tuberculosis has been extensively studied and is of substantial interest because cholesterol is a required nutrient for the survival and persistence of this pathogen in the host (8, 9). In light of the importance of the Ksh system in biotechnology, pathogenesis, and medicine development, the biochemical properties of Ksh subsequently have been characterized in detail using a reconstituted Ksh system from R. rhodochrous DSM 43269 and crystal structure analysis of KshA from M. tuberculosis H37Rv (5). The structure of KshA clearly showed an extended substrate-binding pocket and an active-site-accessing funnel that was consistent with the binding requirement of large steroid substrates (3). The physiological roles of Ksh in these two strains were subsequently investigated in detail (10, 11). In M. tuberculosis H37Rv, only one Ksh is encoded by Rv3526 (kshA) and Rv3571 (kshB). The deletion of either gene resulted in rapid death of the pathogen in the host, indicating that the Ksh system is an essential factor for the infectivity of M. tuberculosis H37Rv (10). In R. rhodochrous DSM 43269, five KshA homologues (KshA1 to KshA5) have been identified (12), and their enzymatic properties have been characterized. Each Ksh system displays unique steroid induction patterns and substrate ranges, indicating an interesting multiplicity of the Ksh system in certain microorganisms that employ dynamic and finely tuned steroid catabolism (11).

Although detailed information has been obtained about Ksh in some microorganisms, such as R. rhodochrous DSM 43269, attempts to modify the Ksh system toward engineered microorganisms that can produce desirable sterol metabolites in high purity and productivity have rarely been successful. Nevertheless, genetic modification of the Ksh system in some sterol-digesting microorganisms has shown remarkable biotechnological significance in the development of valuable steroid producers (13). For example, the five-KshA-null mutant of R. rhodochrous DSM 43269 could transform some sterols into ADD and 3-oxo-23,24-bisnorchol-1,4-dien-22-oic acid (1,4-BNC) at molar ratios of 1% to 7% and 50% to 80%, respectively, depending on the substrate that was used (12). Furthermore, a blocked kshA mutant of M. smegmatis mc2155 that was isolated by transposon mutagenesis was able to produce both AD and ADD from sitosterol (13). These examples demonstrate that the modification of the Ksh system can result in the production of valuable C19 and C22 metabolites from sterols, which is consistent with theoretical speculation. These studies, however, also highlighted the multiplicity and variability of Ksh systems in different microorganisms and illustrated the complexity of Ksh systems with respect to the overall catabolism of sterols. Consequently, a systematic metabolic engineering strategy is required to further develop desirable steroid-producing microorganisms by genetically modifying key reactions in the sterol catabolism, including the Ksh system.

Previously, we characterized and engineered genes encoding key enzymes that participate in sterol catabolism within a fast-growing mycobacterial strain, Mycobacterium neoaurum ATCC 25795 (designated Mn25795 in this study). The engineered genes included the product-determining KstD genes (kstD genes) and two productivity-related cholesterol oxidase genes (choM genes) (1, 4). These experiments successfully developed AD- and ADD-producing strains that exhibit high product purity and productivity (1, 14). In this study, we identified and characterized the Ksh system in Mn25795 to investigate its physiological role in determining the conversion of intermediates during the catabolism of sterols. We then modified the Ksh system using a combined metabolic engineering strategy involving kstD genes and hsd4A (encoding Hsd4A, a dual-function enzyme exhibiting NAD+-dependent 17β-hydroxysteroid dehydrogenase and β-hydroxy-acyl-coenzyme A [CoA] dehydrogenase activity) to develop 9α-OHAD- and ADD-producing strains.

RESULTS

The Ksh system of Mn25795.

KshA is the functional component of Ksh that performs the specific 9α-hydroxylation of 3-ketosteroids. Among characterized KshAs, the two highly conserved domains are the Rieske [2Fe-2S] domain and the nonheme Fe(II)-binding domain (3). Genome-wide sequencing and investigation of the Mn25795 genome revealed two putative KshA-encoding homologues that were designated kshA1N and kshA2N. kshA1N encodes a protein (KshA1N) with 395 amino acids and an estimated molecular mass of 45.2 kDa that shares high amino acid sequence identity with characterized KshAs from M. tuberculosis H37Rv (KshAH37Rv, 84%) and M. smegmatis mc2155 (KshAMS, 89%). kshA2N encodes a protein (KshA2N) with 380 amino acids and an estimated molecular mass of 43.0 kDa that displays high amino acid sequence identity with a putative KshA from M. abscessus (81%) but otherwise exhibits a low sequence identity (<65%) with characterized KshAs. KshA2N and KshA1N share only 62% sequence identity, but the Rieske type [2Fe-2S]R domain (C-X-H-X15–17-C-X2-H) (15) and the mononuclear, nonheme Fe(II) binding motif (D/E-X3-D-X2-H-X4-H) (16), which are typical of KshA, are conserved in both homologues.

kshA1N is located in the sterol catabolism gene cluster, and its promoter region has a palindrome sequence (TnnAACnnGTTnnA, where “n” represents any base), a typical recognition sequence of the highly conserved transcriptional repressor (KstR) of sterol catabolism in mycobacteria (17). The genetic organization of kshA1N and its adjacent genes shares high conservation with its orthologs in M. vanbaalenii PYR-1, M. smegmatis mc2155, and M. tuberculosis H37Rv (Fig. 2a). Consequently, these properties indicated that kshA1N plays a key role in sterol catabolism. Unlike kshA1N, kshA2N was not arranged in the recognized sterol catabolism gene cluster, and no identifiable binding site for KstR was identified in its promoter region, which indicates that kshA2N does not function in a KstR-dependent regulatory mode (18). Nevertheless, the genetic organization of kshA2N and its adjacent genes also shows high conservation in some mycobacteria. Intriguingly, the conserved adjacent genes of kshA2N include two characterized genes, kstD2 and kstD3, which are redundant genes of kstD1 and have been confirmed to play important roles in sterol catabolism (4). Although the function of this gene cluster is unknown, it appears to be related to the degradation of steroids. These observations indicate that kshA2N also is important in the catabolism of sterols.

FIG 2.

FIG 2

Molecular characterization of KshA1N and KshA2N. (a) Schematic description of the locations of kshA1N and kshA2N in the Mn25795 chromosome. Arrow directions indicate the transcriptional direction of genes. Orthologous genes are indicated by gray-shaded regions, kshA1N and kshA2N are shown in green, and kstD genes are shown in blue. Intergenic areas containing the conserved palindromic motif for KstR1 binding in Mn25795 are labeled with asterisks. The genes from M. tuberculosis H37Rv that we predicted to be essential for survival in the host are indicated by triangles. (b) ClustalW alignment of the partial sequence from the putative β-sheet region located at the entrance of the KshA active site from M. tuberculosis H37Rv and Mn25795 and R. rhodochrous DSM 43269. Highly conserved residues and the four mutation points in the chimeric KshA1N are indicated by filled and open rhombuses, respectively. (c) Three-dimensional model structure of KshAH37Rv and the loop regions of KshAH37Rv, KshA1N, KshA2N, KshA1DSM43269, and KshA5DSM43269. Green indicates the putative β-sheet domain; red and blue indicate the first and second α-helix domains, respectively; the orange sphere shows the nonheme ferrous iron; and the orange and yellow cluster spheres indicate the Rieske [2Fe-2S] cluster. The loop region of KshAH37Rv is indicated with a box. The loop regions from KshA1N, KshA2N, KshA1DSM43269, and KshA5DSM43269 were homologously modeled with KshA1H37Rv as a template and are shown as enlarged stick representations.

On the basis of the structure of KshAH37Rv and the structural comparison of substrate preference between KshA1 and KshA5 from R. rhodochrous DSM 43269 (Fig. 2b), a highly variable 19-amino-acid-residue loop region was confirmed to be the essential determinant of substrate preference, and it is located at the entrance of the active site of KshA (6, 19). Alignment indicated that the variable loop region of KshA1N was highly consistent with that of KshAH37Rv, except for a few amino acids, including Asp215 and Ser220 in KshAH37Rv, which corresponded to Asn213 and Thr218 in KshA1N, respectively. Nevertheless, the corresponding loop of KshA2N displays substantial differences from KshAH37Rv, which implies enzymatic differences between KshA1N and KshA2N. To further characterize their properties, the specific loop regions of KshA1N (Glu196 to Val260) and KshA2N (Glu192 to Val258) were homologically modeled using KshAH37Rv as a template. We predicted the loop structure of KshA1N to be significantly different from that of KshA2N and others (Fig. 2c), which further suggests enzymatic differences between KshA1N and KshA2N.

KshB is the auxiliary component of Ksh that provides electrons to KshA. Only one copy of kshB (kshBN) is in the genome, which contains an open reading frame (ORF) of 1,056 bp encoding 351 amino acids with an estimated molecular mass of 37.8 kDa. The protein is a typical class IA monooxygenase reductase harboring three highly conserved domains: a flavin-binding domain (RCYSL, from residues 65 to 69), an NAD binding domain (GSGITP, from residues 129 to 134), and a 2Fe-2S iron-sulfur cluster binding domain (CX4CX2CX29C, from residues 298 to 336). BLAST analysis demonstrated that KshBN shares high amino acid sequence identity with homologues from M. smegmatis mc2155 (KshBMS, MSMEG_5925, 79%) and M. tuberculosis H37Rv (KshBH37Rv, Rv3571, 72%). Like kshBMS and kshBH37Rv, kshBN is located in the gene cluster encoding sterol catabolism-related genes, but it is located several kilobases away from kshA1N. Interestingly, KshBN, along with KshA1N and KshA2N, exhibited more than 99% sequence identity with the putative KshB and two putative KshAs from Mycobacterium sp. strain VKM Ac-1815D (18). Phylogenetic analysis of 16S rRNA genes indicated that Mycobacterium sp. strain VKM Ac-1815D is identical to Mn25795, with 100% sequence similarity.

Determination of the roles of kshA1N and kshA2N in sterol catabolism.

To distinguish the roles of kshA1N and kshA2N in sterol catabolism, we constructed in-frame deletion mutants of kshA1N (MutkshA1N), kshA2N (MutkshA2N), and both together (MutKshA1&2N), in addition to their functional complements (ComkshA1N, ComkshA2N, and ComkshA1&2N, respectively), in Mn25795. The deletion of kshA1N significantly inhibited the growth of Mn25795 in minimum medium (MM) with cholesterol as the sole carbon source (Fig. 3a). Moreover, the inhibitory effect was clearly observed in MutkshA1N grown in MM with two carbon sources (i.e., cholesterol and glycerol), demonstrating the critical role of kshA1N in the catabolism of cholesterol. In contrast, kshA2N appeared to have no obvious effect on cholesterol catabolism, as the single deletion of kshA2N did not disturb the growth of Mn25795 using cholesterol or cholesterol-glycerol as carbon sources. To exclude the shielding effect from the high activity of KshA1N, the role of kshA2N was again analyzed after the deletion of kshA2N and kshA1N, which showed that the double deletion of kshA1N and kshA2N resulted in even worse growth on cholesterol. Therefore, kshA2N may provide functional redundancy for kshA1N.

FIG 3.

FIG 3

Determination of the role of kshA1N and kshA2N in Mn25795. (a) Growth curves of Mn25795 and its derivative strains using cholesterol (a) or cholesterol-glycerol (b) as carbon sources, respectively. Symbols: Mn25795, ■; MutkshA1N, △; MutkshA2N, ▽; MutkshA1&2N, ♢; ComkshA1N, ▲; ComkshA1&2N, ◆; Com-blank, •. The concentration of cholesterol was 2 g liter−1. The concentration of glycerol was the same as that for MYC/01. (c) High-performance liquid chromatography (HPLC) analyses of the metabolites resultant from the metabolism of 2 g liter−1 cholesterol by MutkshA1N in cholesterol-glycerol medium at 120 h. (d) Transcriptional changes in kshA1N and kshA2N in glycerol medium with the addition of 2.0 mM cholesterol and 1.5 mM AD, respectively. Data are shown as averages from triplicate experiments ± standard errors.

Two hypothetical mechanisms may lead to growth impairment caused by KshA1N inactivation. One mechanism occurs by blocking the degradation of the cholesterol steroid nucleus, such that no efficient carbon source would be available to support growth. The other mechanism involves the accumulation of some cholesterol metabolites, such as AD and ADD, that are toxic to respiration, viability, and duplication in microorganisms (20). To assess whether these putative metabolic intermediates exhibited strong toxicity and inhibited the growth of MutkshA1N, we added glycerol as a supplemental carbon source to the MM containing cholesterol to evaluate growth differences. Growth of MutkshA1N in the cholesterol-glycerol medium resulted in considerable accumulation of metabolites, including AD and ADD, whereas growth of MutkshA1N only slightly changed when grown in glycerol medium and the cholesterol-glycerol medium (Fig. 3b and c). These results indicate that cholesterol metabolism intermediates had no significant inhibitory effects on the growth of MutkshA1N. Therefore, the significant inhibition of growth caused by the deletion of kshA1N could mainly be attributed to the lack of an appropriate carbon source to support growth. Compared with MutkshA1N, the growth of MutkshA2N in cholesterol-glycerol medium showed no obvious accumulation of cholesterol metabolites. These results clearly demonstrated that kshA1N is the main kshA gene responsible for cholesterol metabolism.

The potential role of kshA2N in sterol catabolism was thus unclear, as kshA2N is not a member of the recognized gene cluster encoding sterol catabolism enzymes, and these results did not reveal its definite function in sterol metabolism. Therefore, to resolve this uncertainty and to distinguish functional differences between kshA1N and kshA2N, we measured transcription levels of kshA1N and kshA2N by reverse transcription-quantitative PCR (RT-qPCR) with the addition of cholesterol and AD to culture medium and with glycerol as the control. Both cholesterol and AD significantly induced the transcription of kshA1N and kshA2N, suggesting that kshA2N is also involved in sterol metabolism, similar to kshA1N (Fig. 3d). Cholesterol addition induced an approximately 2.0-fold increase in ksh2N transcription and an approximately 4.7-fold increase in kshA1N transcription, confirming that kshA1N was more important than kshA2N in sterol catabolism. Moreover, both kshA1N (1.7-fold) and kshA2N (1.6-fold) were induced by AD, which is a common C19 steroid intermediate in sterol catabolism. Furthermore, the induction effect of cholesterol was greater than that of AD for kshA1N and also higher than that of AD for kshA2N but to a lesser extent, indicating that AD is a poor inducer of kshA1N and kshA2N.

Enzymatic characterization of KshA1N and KshA2N in vitro.

To determine whether KshBN could form the active Ksh enzyme complex with KshA1N and KshA2N, the Ksh enzymes were reconstituted from separately expressed and purified KshBN and KshAN (KshA1N or KshA2N). However, no Ksh activity was detected, though the purified KshBN maintained reductase activity of 6.3 μmol min−1 mg−1 in a 2,6-dichlorophenolindophenol (DCPIP) assay. As suggested by Petrusma et al., the protein-protein interactions between KshA and KshB subunits are critically important for maintaining Ksh activity (5). Therefore, the copurification of KshAN and KshBN was conducted, and the active Ksh system was successfully reconstituted (Fig. 4a and b). To ensure the full activity of KshAN in the reconstructed Ksh system, an additional amount of purified KshBN was used to supplement the copurified KshAN and KshBN to achieve a KshAN-to-KshBN molar ratio of approximately 1:2. Under these conditions, we analyzed the enzymatic properties of the Ksh enzymes consisting of KshA1N or KshA2N copurified with KshB (Tables 1 and 2).

FIG 4.

FIG 4

SDS-PAGE analysis of E. coli expression and copurification of KshAN and KshBN. (a) KshA1N and KshBN. (b) KshA2N and KshBN. Lane 1, standard marker protein; 2, cell extracts from BL21-kshAN; 3, cell extracts from BL21-kshBN; 4, mixture of cell extracts of BL21-kshAN and BL21-kshBN in a KshAN:kshBN molar ratio of approximately 1:2; 5, the copurified fraction of KshAN and KshBN with 150 mM imidazole.

TABLE 1.

Steroid substrate profiles of KshA1N and KshA2N

Steroid substratea KshA1N
KshA2N
Enzyme activitya (nmol min−1 mg−1) Relative activityb (%) Enzyme activity (nmol min−1 mg−1) Relative activity (%)
4-Androstene-3,17-dione (AD) 2.3 × 102 ± 21 × 102 100 1.2 × 102 ± 15 × 102 100
4-Androstene-17β-ol-3-one (testosterone) 3.1 × 102 ± 17 × 102 135 90 ± 10 75
1,4-Androstadiene-3,17-dione (ADD) 6.5 × 102 ± 31 × 102 282 1.5 × 102 ± 29 × 102 125
1,4-Androstadiene-17β-ol-3-one (boldenone) 5.8 × 102 ± 47 × 102 252 1.3 × 102 ± 27 × 102 108
1-(5α)-Androstene-17β-ol-3-one 1.8 × 102 ± 39 × 102 78 23 ± 7 19
1-(5α)-Androstene-3,17-dione 2.1 × 102 ± 33 × 102 91 27 ± 9 23
5α-Androstane-3,17-dione 52 ± 11 23 NDc ND
5β-Androstane-3,17-dione 75 ± 12 33 ND ND
5α-Androstane-17β-ol-3-one (androstanolone) 42 ± 15 18 ND ND
5α-Androstane-3α-ol-17-one ND ND ND ND
5α-Androstane-3β-ol-17-one (epiandrosterone) ND ND ND ND
5-Androstene-3β-ol-17-one (dehydroepiandrosterone) ND ND ND ND
9α-Hydroxy-4-androstene-3,17-dione (9-OHAD) ND ND ND ND
4-Androstene-3,17-dione-19-ol (19-OHAD) 23 ± 16 10 ND ND
1,3,5(10)-Estratrien-3-ol-17-one ND ND ND ND
4-Pregnene-3,20-dione (progesterone) 7.7 × 102 ± 49 × 102 335 1.8 × 102 ± 21 × 102 150
11β-Hydrocortisone 1.2 × 102 ± 40 × 102 52 38 ± 11 32
23,24-Bisnorchol-1,4-diene-22-oic acid (1,4-BNC) 9.8 × 102 ± 55 × 102 426 2.1 × 102 ± 38 × 102 175
4-Cholestene-3-oned (cholesterone) 45 ± 7 21 ND ND
5-Cholestene-3β-old (cholesterol) ND ND ND ND
a

Specific Ksh activity levels toward substrates were analyzed at 200 μM and represented as means ± standard errors from three independent experiments.

b

Relative activity signifies the percentage of the activity toward substrates relative to that of AD, which was set at 100%.

c

ND, no detectable activity toward specific steroid.

d

Steroid concentration of 25 μM due to the low steroid solubility.

TABLE 2.

Relative initial activities of wild-type and mutated KshA1N with a range of steroid substrates

Steroid substrate Relative initial activity of Ksh (%)a
A1N A1V207T A1P210E A1D225R A1N238D
4-Androstene-3,17-dione (AD) 100 100 100 100 100
4-Androstene-17β-ol-3-one (testosterone) 135 ± 10 161 ± 30 154 ± 20 133 ± 37 139 ± 24
1,4-Androstadiene-3,17-dione (ADD) 282 ± 40 289 ± 38 275 ± 47 178 ± 29 252 ± 50
1,4-Androstadiene-17β-ol-3-one (boldenone) 252 ± 31 258 ± 33 255 ± 51 174 ± 18 267 ± 29
1-(5α)-Androstene-3,17-dione 91 ± 19 112 ± 20 100 ± 22 63 ± 18 95 ± 27
5α-Androstane-3,17-dione 23 ± 4 32 ± 16 32 ± 6 NDc 26 ± 9
5β-Androstane-3,17-dione 33 ± 6 42 ± 18 41 ± 12 15 ± 5 52 ± 19
5α-Androstane-17β-ol-3-one (androstanolone) 18 ± 3 35 ± 11 13 ± 7 ND 25 ± 9
17α-Methyl-1-(5α)-androstane-17β-ol-3-one 78 ± 8 92 ± 21 82 ± 10 32 ± 13 73 ± 32
4-Androstene-3,17-dione-19-ol 10 ± 2 12 ± 4 16 ± 2 ND ND
4-Pregnene-3,20-dione (progesterone) 335 ± 38 430 ± 55 323 ± 40 215 ± 61 320 ± 40
11β-Hydrocortisone 52 ± 9 82 ± 26 42 ± 11 22 ± 8 57 ± 19
23,24-Bisnorchol-4-ene-22-oic acid (4-BNC) 426 ± 52 499 ± 58 447 ± 65 260 ± 43 440 ± 52
4-Cholestene-3-oneb (cholestenone) 21 ± 2 56 ± 17 30 ± 9 ND 25 ± 12
a

Relative activity signifies the percentage of activity toward substrates at 200 μM relative to that of AD, which was set at 100%. Errors were calculated as standard errors of the means (from three experiments).

b

Steroid concentration of 25 μM due to the low solubility of cholesterone.

c

ND, no detectable activity toward specific steroid.

Both KshAN homologues showed a broad preference for steroid substrates with no substantial differences, although the specific activity of KshA1N was higher than that of KshA2N for suitable substrates (Table 1). Clearly, steroids with 4-ene-3-one or 1-ene-3-one structures were the proper substrates for both KshAN homologues, whereas steroids with 3-alcohol structures or aromatic A rings were not. Moreover, 1,4-diene-3-one structures, such as those of ADD and boldenone, were the more suitable substrates for both KshAN homologues than 4-ene-3-one substrates, such as AD and testosterone. These substrate preferences indicate that the 3-keto-4-ene-conjugated moiety was the required structure for proper substrates of both KshANs. The thing of note, however, was that KshA1N also showed a small amount of activity with some steroids with 3-keto structures and without double bonds on the A ring, including androstanolone and androstane-3,17-dione.

Interestingly, both KshAN homologues displayed the highest activities toward 23,24-bisnorchol-4-ene-22-oic acid (4-BNC) among the tested substrates, followed by progesterone, another steroid with a short C-17 side chain, which were much higher than their activities toward the C19 steroids. These results indicated that the optimal substrates of KshAN should also contain C-17 side chains. Cholesterone, a sterol intermediate with a full C-17 side chain, was not a good substrate for both KshAN homologues. Thus, only certain intermediates in the sterol catabolism pathway with a truncated (partially degraded) C-17 side chain, such as 1,4-HBC and its derivatives or precursors, were optimal substrates for KshA1N and KshA2N rather than C19 steroids. This clearly demonstrates that the cleavage of the steroid nucleus should occur concurrently with the cleavage of the C-17 side chain rather than afterward, which is consistent with what has been proposed for KshAB in M. tuberculosis (21).

To determine substrate affinities, we performed kinetic studies to determine the affinity constants of KshA1N and KshA2N. The enzyme activities of KshA1N and KshA2N toward different concentrations of substrates are given in Tables S3 and S4 in the supplemental material. Enzyme activity of KshA1N toward most of the tested substrates showed a Km value below 10 μM, except for androstane structures, cholesterone, and 4-androstene-3,17-dione-19-ol, indicating a desirable enzyme-substrate binding relation. The substrate affinities for KshA2N had a Km value of 33 ± 8 μM for 1-(5α)-androstene-17β-ol-3-one, 25 ± 6 μM for 1-(5α)-androstene-3,17-dione, and 13 ± 3 μM for 11β-hydrocortisone; the remaining substrates that were tested had Km values of <10 μM (Table 1).

Modification of KshA1N.

Given the pivotal role of Ksh in the metabolism of sterols, its activity must be modified to develop engineered strains that produce steroidal metabolites with sufficient yield. In some cases, such as in the development of 9-OHAD-producing strains, sufficient activity of Ksh is necessary to reduce the generation of side products (e.g., AD). Therefore, this study further investigated the modification of KshA1N.

Previous studies have demonstrated that KshA homologues have similar tertiary structures with a variable β-sheet, including 58 amino acids in the substrate-binding pocket. In addition, a highly variable loop region of the β-sheet at the entrance of the active site contributes to substrate preference (Fig. 2b and c) (19, 22). Petrusma et al. demonstrated that the β-sheet could be exchanged between two KshA homologues, and the resulting chimeric enzymes exhibited substrate preferences similar to those of the donor enzymes (19). Hence, the β-sheet and loop regions of KshA1 and KshA2 were exchanged in this study to evaluate whether the enzymatic properties could be modified, which resulted in the chimeric proteins KshA1A2β, KshA2A1β, KshA1A2loop, and KshA2A1loop. Unfortunately, the four chimeras exhibited no enzymatic activity, indicating that the β-sheet and loop structures of KshA1 and KshA2 are not exchangeable.

Consequently, we attempted a strategy of point mutation of the β-sheet to modify the activity of KshA. On the basis of previous reports and alignment of the KshAN β-sheets with well-characterized KshA counterparts from M. tuberculosis H37Rv and R. rhodochrous DSM 43269 (Fig. 2b), we selected four putatively crucial but nonconserved amino acids in the β-sheet of KshA1N (two in the loop region) as the mutation sites (Fig. 2b): V207, P210, D225, and N238. According to the mutation of Q209T in the KshA1A5 chimera that was reported previously (19), V207 was replaced by T207, resulting in a V207T mutant. We conducted the other three point mutations (P210E, N238D, and D225R) based on the differences in amino acids between Mycobacterium and Rhodococcus in the alignment analysis (Fig. 2b). The point mutations P210E (activity toward AD, 2.1 × 102 ± 18 × 102 nmol min−1 mg−1), N238D (activity toward AD, 1.9 × 102 ± 25 × 102 nmol min−1 mg−1), and D225R (activity toward AD, 1.6 × 102 ± 15 × 102 nmol min−1 mg−1) displayed no obvious improvement in activity relative to wild-type KshA1N (Table 2). The KshA1V207T construct showed elevated Ksh activity (AD, 2.4 × 102 ± 23 × 102 nmol min−1 mg−1) relative to the wild-type enzyme, however, and this was particularly evident with increases of 17.1% and 28.4% toward 4-BNC and progesterone, respectively (Table 2). As described earlier, sterol intermediates with a truncated side chain are the proper physiological substrates of Ksh. Thus, the significantly enhanced activity of KshA1V207T toward 4-BNC indicated an apparently improved activity of KshA1V207T relative to KshA1N. Therefore, we employed KshA1V207T in further modification of Mn25795.

Development of an ADD-producing strain.

ADD is a valuable C19 metabolite that can be used as a precursor for the industrial synthesis of 19-nor-steroids, estrogens, and some adrenocortical hormones. As the abrogation of Ksh activities can theoretically block the continuous degradation of ADD and lead to ADD accumulation during sterol catabolism, we attempted to develop an ADD-producing strain of Mn25795 via inactivation of KshANs. As indicated, the kshA1N-kshA2N double-deletion strain MutkshA1NA2N (termed strain I) could transform phytosterols with ADD constituting the major metabolic product. We added 2 g liter−1 phytosterols to fermentation media with strain I, followed by incubation in a shake flask for 168 h. The resultant conversion rate of phytosterols was nearly 95%, and the ADD yield reached 0.46 g liter−1, which represents a 33.20% molar yield. The strain also produced other by-products in addition to ADD, including a 20.6% yield (0.29 g liter−1) of AD, a 15.1% (0.21 g liter−1) yield of DHT, and a 4.4% yield (0.06 g liter−1) of T. Thus, the construct MutkshA1NA2N was not a great ADD producer and could be further improved by reducing the production of these by-products. Considering the role of KstD in transforming AD to ADD (Fig. 1b), the undesired production of AD was attributed to insufficient activity of KstD. Previously, three KstD homologues, and their physiological roles, were characterized in Mn25795, with KstD1 representing the primarily active homologue (4).

Following these observations, kstD1 was overexpressed in MutkshA1NA2N, resulting in the strain MutkshA1NA2N-p261kstD1, which was termed strain II. After incubation for 168 h with 2 g liter−1 phytosterol additions, strain II gave a markedly enhanced yield of ADD (56.3% molar yield, 0.78 g liter−1) with a significantly decreased yield of AD (5.4% molar yield, 0.08 g liter−1) (Table 3) compared with those of strain I. Moreover, no detectable DHT and T were produced. The ADD yield ratio among the metabolic products of strain II reached 91.2%, representing more than a 2.0-fold increase (45.3%) over strain I (Fig. 5b). These results indicate the necessity of combined modification of kshA and kstD genes in the development of ADD production strains. To evaluate the potential for industrial application of strain II, we employed a resting cell biotransformation system in a shake flask with 20 g liter−1 phytosterols. Surprisingly, we obtained only a 38.2% molar yield (5.3 g liter−1) of ADD, along with abundant 1,4-HBC (31.0% yield, 4.3 g liter−1) (Fig. 5c). Intriguingly, these results differed from previous results, yielding the new by-product 1,4-HBC and indicating that the sterol transformation conditions have an important effect on the resultant by-products.

TABLE 3.

Profiles of phytosterol transformation by ADD-producing strains for 168 h

Strain and concn of substratea (g liter−1) Major product Concn (g liter−1) Molar yieldb (%) Yield ratio of ADDc (%)
I
    2 ADD 0.46 ± 0.03 33.2 ± 1.8 45.3 ± 0.5
AD 0.29 ± 0.04 20.6 ± 2.7
DHT 0.21 ± 0.01 15.1 ± 1.0
T 0.06 ± 0.01 4.4 ± 1.1
II
    2 ADD 0.78 ± 0.05 56.3 ± 3.2 91.2 ± 0.3
AD 0.08 ± 0.02 5.4 ± 0.5
    20 (resting cells) ADD 5.31 ± 0.25 38.2 ± 1.8 55.2 ± 0.6
1,4-HBC 4.29 ± 0.20 31.0 ± 1.5
    20 (growing cells) ADD 7.78 ± 0.15 55.4 ± 1.2 89.1 ± 0.2
AD 0.60 ± 0.03 4.3 ± 0.2
1,4-HBC 0.41 ± 0.01 2.5 ± 0.1
III
    20 (resting cells) ADD 7.51 ± 0.31 53.8 ± 2.1 72.6 ± 0.4
1,4-HBC 2.82 ± 0.29 20.3 ± 1.7
    20 (growing cells) ADD 9.36 ± 0.25 66.7 ± 1.6 92.6 ± 0.3
AD 0.74 ± 0.05 5.3 ± 0.3
a

The conversion of 2 g liter−1 phytosterols was performed in shake flasks using growing cells, and the conversion of 20 g liter−1 phytosterols was performed in a 7-liter bioreactor using resting cells or growing cells. Data are shown as the averages from triplicate experiments.

b

Amounts of major metabolites and their molar yields were determined from three independent transformations and expressed as the actual range from the least to the topmost.

c

The yield ratio of ADD was obtained by calculating the molar yield of ADD among all the major products.

FIG 5.

FIG 5

ADD synthesis metabolic pathway and metabolite analysis of constructed ADD-producing strains. (a) The ADD synthesis metabolic pathway from β-sitosterol in Mn25795. Red arrows indicate strengthened steps, and the crossed arrows represent blocked steps. The structure of the desired product ADD is also indicated in red. (b) HPLC comparison of the products derived from transformation of 2 g liter−1 phytosterol by strains I and II in the growing cell system at the shake flask level. (c) HPLC comparison of the products derived from transformation of 20 g liter−1 phytosterols by strains II and III in the resting cell system at the shake flask level.

On the basis of our previous research (23), the accumulation of 1,4-HBC during the sterol catabolism can be attributed to the insufficient activity of a key enzyme, Hsd4A, that is involved in the degradation of the sterol side chain (Fig. 5a). Hsd4A is a dual-function enzyme in sterol catabolism, exhibiting NAD+-dependent 17β-hydroxysteroid dehydrogenase and β-hydroxy-acyl-CoA dehydrogenase activity. The latter activity was considered to be responsible for the conversion of 22-hydroxy-3-oxo-25,26-bisnorchol-4-en-24-oyl CoA (22HOBNC-CoA) to 3,22-dioxo-25,26-bisnorchol-4-ene-24-oyl CoA (22OBNC-CoA), and the deletion of Hsd4A thus could result in the production of 1,4-HBC or its derivatives (23). Therefore, we further modified strain II by increasing the expression of Hsd4A to reduce the yield of 4-HBC, thereby resulting in MutkshA1NA2N-p261kstD1-p60hsd4A, termed strain III. After incubation with 20 g liter−1 phytosterols in the resting cell system for 168 h, strain III gave a 53.8% molar yield (7.5 g liter−1) of ADD and maintained a 20.3% molar yield (2.8 g liter−1) of 1,4-HBC (Table 3). Compared with that of strain II, the ADD yield ratio of strain III increased from approximately 55.2% to 72.6%, indicating the effectiveness of hsd4A overexpression toward the production of ADD. Consequently, we further evaluated the productive properties of strains II and III in the growing cell system. The results indicated that strain II can ferment 20 g liter−1 phytosterols in 168 h to yield 55.4% ADD (7.78 g liter−1), 4.3% AD (0.60 g liter−1), and 2.5% 1,4-HBC (0.41 g liter−1). In contrast to the results from the resting cell system, the 1,4-HBC yield was significantly reduced from a 31.0% molar yield to 2.5%, although a 4.3% molar yield of AD still occurred, and the yield ratio of ADD was greatly enhanced, from 55.2% to 89.1%.

These results highlight the importance of transformation growth conditions in determining the composition of metabolic products from the conversion of phytosterols by ADD-producing strains. We also converted 20 g liter−1 phytosterols using strain III in the growing cell system and obtained a 66.7% molar yield of ADD (9.36 g liter−1), the yield ratio of which reached 92.5%, along with a 5.3% molar yield of AD (0.74 g liter−1) and no measurable 1,4-HBC yield (Table 3). These results demonstrate that strain III is a good ADD-producing strain and that the growing cell system is more suitable than the resting cell transformation system for the production of ADD from phytosterol conversion.

Development of a 9-OHAD-producing strain.

9-OHAD is a valuable C19 metabolite and is a good precursor for the industrial production of adrenocortical hormones. The construction of a 9-OHAD-producing strain also involves the modification of Ksh activity. In our previous work (4), M. neoaurum NwIB-yV was constructed via the overexpression of KshA1N in strain M. neoaurum MutMN-kstD1&2&3, which is designated strain IV in this study. After 120 h of incubation with 2 g liter−1 of phytosterols in a shake flask, strain IV exhibited a conversion rate of >95% over 120 h. The conversion rate of this 9-OHAD-producing strain was higher than that of ADD-producing strains, for which we obtained a 74.8% molar yield (1.10 g liter−1) of 9-OHAD, along with a 3.4% yield (0.05 g liter−1) of AD (Table 4). The 9-OHAD yield ratio of strain IV was more than 95%, indicating that strain IV was a good 9-OHAD producer. Subsequently, we investigated its industrial application potential in the resting cell system with 20 g liter−1 phytosterols. However, the 9-OHAD yield ratio of strain IV decreased significantly from 95.7% to 89.0%, with a 68.3% molar yield (10.04 g liter−1) of 9-OHAD and an 8.4% molar yield (1.24 g liter−1) of AD. These results indicate that in the presence of high concentrations of phytosterols, the activity of Ksh is insufficient and should be further enhanced to improve the yield ratio of 9-OHAD. The transcript levels of kshA1N in strain IV were 1.6-fold and 2.0-fold higher than that of strain MutMN-kstD1&2&3 after 72 h and 120 h of phytosterol transformation, respectively (Fig. 6b).

TABLE 4.

Profiles of phytosterol transformation by 9-OHAD-producing strains for 120 h

Strain and concn of substratea (g liter−1) Major product Concnb (g liter−1) Molar yieldc (%) Yield ratio of 9-OHAD (%)
IV
    2 9-OHAD 1.10 ± 0.05 74.8 ± 3.1 95.7 ± 0.2
AD 0.05 ± 0.02 3.4 ± 1.0
    20 9-OHAD 10.04 ± 0.22 68.3 ± 1.5 89.0 ± 0.5
AD 1.24 ± 0.11 8.4 ± 0.5
V
    2 9-OHAD 1.32 ± 0.03 90.2 ± 1.9 97.6 ± 0.3
AD 0.03 ± 0.01 2.2 ± 0.3
    20 9-OHAD 11.71 ± 0.15 79.8 ± 1.0 94.2 ± 0.4
AD 0.71 ± 0.10 4.9 ± 0.6
a

The conversion of 2 g liter−1 phytosterols was performed in shake flasks using growing cells, and the conversion of 20 g liter−1 phytosterols was performed in a 7-liter bioreactor using resting cells. Data are shown as the averages from triplicate experiments.

b

Amounts of major metabolites and their molar yields were determined from three independent transformations and are expressed as the actual range from the least to the topmost.

c

The yield ratio of 9-OHAD was obtained by calculating the molar yield of 9-OHAD among all the products.

FIG 6.

FIG 6

Comparative analyses of 9-OHAD-producing strains. (a) The role of Ksh in the production of 9-OHAD. (b) Transcription profiles of kshA1N in strains IV and V at 72 h (gray bars) and 120 h (dark bars) during the metabolism of 2.0 mM cholesterol. The fold change values indicate mRNA levels in strains IV and V relative to that of MutMN-kstD1&2&3. The data represent averages from triplicate experiments, and the error bars indicate ± standard deviations. (c) HPLC comparison of metabolites from the transformation of 2 g liter−1 phytosterols by strains IV and V in the growing cell system at the shake flask level. (d) HPLC comparison of the metabolites from the transformation of 20 g liter−1 phytosterols by strains IV and V in the resting cell system at the shake flask level.

These observations suggest that KshA1N already is well overexpressed in strain IV, and increasing the expression level of KshA1 further likely will not improve the conversion of AD to 9-OHAD. Therefore, to enhance activity compared with that of KshA1, we overexpressed the mutant KshA1V207T in MutMN-kstD1&2&3 to construct an improved 9-OHAD-producing strain relative to strain IV, thereby resulting in MutMN-kstD1&2&3-p261-kshA1V207T, termed strain V. As expected, strain V showed improved performance in the resting cell system, converting 20 g liter−1 of phytosterols to 11.7 g liter−1 9-OHAD within 120 h while also producing only 0.71 g liter−1 AD. Compared with strain IV, the 9-OHAD yield ratio of strain V was enhanced from 89.0% to 94.2%. These results clearly show that strain V is a more promising 9-OHAD producer for industrial application than strain IV based on its higher 9-OHAD yield ratio and decreased by-product formation.

DISCUSSION

In the industrial process of phytosterol transformation to produce valuable steroid intermediates, unavoidable generation of by-products often leads to high purification costs and a low recovery rate of the target product, and therefore it is one of the main barriers limiting the successful development of engineered strains for industrial application. Generally, side products arise from specific rate-limiting reactions that often are concurrent and promiscuous. The 3-ketosteroid-Δ1-dehydrogenation and 3-ketosteroid-9α-hydroxylation that are performed by KstDs and Kshs, respectively, are two key reactions that directly determine the yield and yield ratio of products during sterol catabolism. The two reactions proceed concurrently and can cross-react, which involves many intermediates in the catabolic process of sterols.

The complex multiplicity of kshA is a common feature in rhodococci and mycobacteria, and it supports the utilization of a variety of environmental steroids as nutrition sources. For example, there are five kshA genes in R. rhodochrous DSM 43269 and Mycobacterium sp. strain VKM Ac-1817D and seven kshA genes in R. equi (24). In contrast to the multiplicity in these strains, the Ksh system in Mn25795 exhibited moderate redundancy with two KshA homologues: KshA1N and KshA2N. This suggests that the strain is more easily modified for industrial use by mutation screening or genetic engineering. In line with previously reported KstDs and KshAs (4, 11), KshA1N and KshA2N clearly exhibited wide and differentiated substrate preferences in this study, which suggests promiscuous reaction orders are carried out by Kshs with metabolic intermediates of phytosterols. Among the tested intermediates, both KstDs and KshAs showed greater activity toward the metabolic intermediates with truncated side chains, most notably BNC-CoA, 1,4-BNC-CoA, or 9α-OH-BNC-CoA (4), instead of the other metabolites, including C19 metabolites AD, ADD, and 9-OHAD. For strains engineered to excessively accumulate certain C19 metabolites, some precursors of the expected metabolite, which is not the optimal substrate for KstDs and KshAs, will be inevitably and excessively produced because of the concurrent and promiscuous reactions in the catabolic process of sterols. As most of the precursors are not the optimal substrates of KstDs and KshAs, these excess poor substrates could easily result in the generation of by-products resulting from the competitiveness effect, when KstD and KshA expression does not increase accordingly. This process is not a rare event, as some side products often are generated in industrial sterol biotransformation. For example, AD often is generated as a main by-product in the industrial production of 9-OHAD or ADD (4, 14). The activities of KstDs and Kshs are therefore of substantial interest, and their expression must be accordingly modified in the development of engineered strains for industrial application.

In view of the limiting conditions for cell growth in the shake flasks, such as lowered dissolved oxygen and pH fluctuation, the transformation capacity of growing mycobacterial cells at the shake flask level is generally limited far below their practical transformation capacity in bioreactors. In this study, we employed a resting cell transformation system, which has been developed to rapidly assess the application potential of newly constructed strains in previous studies (25). Interestingly, the resting cell transformation system showed another use as an assessment tool to identify possible rate-limiting steps among the metabolic processes of sterols. The constructed ADD-producing strain II showed a desirable transformation capacity for 20 g liter−1 phytosterols over 168 h when grown in the resting cell transformation system but produced an unexpected by-product, 1,4-HBC. This result is significantly different from the metabolite profile obtained in the growing cell transformation system with 2 g liter−1 phytosterols at the same shake flask level (Table 3). The yield of 1,4-HBC indicated that a rate-limiting step must exist in the degradation of sterol side chains. On the basis of our previous research (23), the rate-limiting step could be attributed to the conversion of 22HOBNC-CoA to 22OBNC-CoA in the degradation of the sterol side chain that is catalyzed by Hsd4A (Fig. 5a). Augmented expression of Hsd4A in strain II greatly decreased the molar yield of 1,4-HBC in the resting cell transformation system and completely inhibited its yield in the growing cell transformation system. These results confirmed that the conversion of 22HOBNC-CoA to 22OBNC-CoA is the rate-limiting step for the production of ADD in strain II. Consequently, hsd4A should be considered for modification in combination with kstD and kshA, which will together be useful in the development of ADD-producing strains via genetic engineering. Obviously, the usefulness of the resting cell transformation system to evaluate possible rate-limiting steps may be attributed to the magnification of effects resulting from specific rate-limiting steps in the resting cell system compared with the growing cell system.

In the phytosterol transformation of strain IV in the resting cell transformation system, the yield ratio of 9-OHAD was significantly decreased because of the increase of AD as a by-product compared with that in the growing cell transformation system, indicating that the activity of Ksh was still insufficient for the transformation of sterols when provided in high concentrations. This was particularly intriguing because the transcript levels of kshA1N in strain IV were 1.6- to 2.0-fold greater than its original levels for the strain without KshAN overexpression. To address this problem, we constructed a mutant, kshA1V207T, with significantly elevated Ksh activity. The mutant KshA1V207T exhibited improved activity in the conversion of phytosterols to 9-OHAD compared with KshA1. Consequently, this development provides an alternative means to improve the industrial potential of 9-OHAD-producing strains.

A summary of the strains that were constructed in this study and the effects of various genotypes on sterol catabolism are shown in the Fig. 7. Overall, this study profiled the properties of 3-ketosteroid 9α-hydroxylases in Mn25795 in detail and investigated specific means to develop ADD-producing and 9-OHAD-producing strains by investigating the production of side products and accordingly modifying kshAN and other key genes that are associated with the yield of by-products. In combination with our previous studies on the characterization and engineering of kstD genes in Mn25795 (14, 23), our present analysis of kshA1N and kshA2N genes provides a relatively complete engineering solution that can be used to develop optimal ADD-producing and 9-OHAD-producing strains for desired industrial applications.

FIG 7.

FIG 7

Metabolic pathways of ADD and 9-OHAD synthesis. (a) Metabolic pathway to produce ADD; (b) metabolic pathway to produce 9-OHAD; (c) description of the strains that were established in this study. ADD and 9-OHAD are denoted in red. The key steps catalyzed by Kshs, KstDs, and Hsd4A were labeled by red, purple, and blue arrows, respectively. The symbol X represents the block of the metabolic pathway, and the thick red arrows, purple arrows, and the blue arrow represent the overexpression of Ksh, KstDs, and Hsd4A in the metabolic pathway. The successive thick arrows represent certain unspecified enzymes.

MATERIALS AND METHODS

Bacterial strains, plasmids, and culture conditions.

All strains, plasmids, and primers used in this study are listed in Table 5 and Table S1 in the supplemental material. Escherichia coli DH5α was employed for plasmid replication, and Escherichia coli BL21(DE3) was used for heterologous expression. Luria-Bertani (LB) medium was used for cultivating E. coli strains, and glycerol medium MYC/01 and glucose medium MYC/02 were used for cultivating M. neoaurum strains as previously reported (1). Before being added to the media, phytosterols were prepared as a 240 mM stock solution, emulsified by Tween 80 via ultrasonication (20 min, 300 W).

TABLE 5.

Primers, plasmids, and strains used in this study

Namea Description Source or reference
Primers
    pETkshA1N-f&r TATCCATGGTGACTACCGAGACAGCCGGC/TATCTCGAGGCTCGGCTGAGCCGGTTCTTT This study
    pETkshA2N-f&r TATCCATGGTGACCGATATCCGCGAGATCG/TATCTCGAGCGCAGTCGAACCGCGACCG This study
    pETkshBN-f&r TGGCCATGGTGACGGAGGAACCGCTCGGCA/TCACTCGAGTTCGTCGTAGGTGACTTCCA This study
    p261kshA1N-f&r CAAGGATCCAGTGACTACCGAGACAGCCGGC/TAGAAGCTTCTACAGGTTTTCCTCGACCTCGA This study
    p261kshA2N-f&r CAAGGATCCAATGACCGATATCCGCGAGATCG/TAGAAGCTTTTACGCAGTCGAACCGCGACCG This study
    p261kstD1-f&r TACGGATCCAGTGTTCTACATGACTGCCCAGGA/GCCAAGCTTTCAGGCCTTTCCAGCGAGATGC This study
    o-p261-f&r TAGGCGAGTGCTAAGAATAACGTTG/TCGTTTTATTTGATGCCTGGCAGT This study
    16SrRNA-f&r CCTATGTTGCCAGCGGGTTATGC/GCGATTACTAGCGACTCCGACTTCA This study
    RTqA1-f&r GCGTCCTACTTCGGTCCGTCGTTCA/GTCGATCCGGGTCTTGTGCTTCCAG This study
    RTqA2-f&r GCCTATCCGACGTACTTCAAGAAC/CACTGCAGGACGAACGAATCATGG This study
    kshA1V207T-f&r TACCTGCACAACACCGGCCGTCCGGAT/ATCCGGACGGCCGGTGTTGTGCAGGTA This study
    kshA1P210E-f&r AACGTCGGCCGTGAGGATGTCAACGAC/GTCGTTGACATCCTCACGGCCGACGTT This study
    kshA1D225R-f&r GAGGCGCACCTGCGTTCCGAGGCGTCC/GGACGCCTCGGAACGCAGGTGCGCCTC This study
    kshA1N238D-f&r TCGTTCATGATCGACTGGCTGCACAAC/GTTGTGCAGCCAGTCGATCATGAACGA This study
Plasmids
    pET-28a(+) E. coli expression vector, Kanr Novagen
    pET28a-kshA1N Expression plasmid pET28a possessing ORF of kshA1N This study
    pET28a-kshA1V207T Expression plasmid pET28a possessing ORF of kshA1V207T This study
    pET28a-kshA1P210E Expression plasmid pET28a possessing orf of kshA1P210E This study
    pET28a-kshA1D225R Expression plasmid pET28a possessing ORF of kshA1D225R This study
    pET28a-kshA1N238D Expression plasmid pET28a possessing ORF of kshA1N238D This study
    pET28a-kshA2N Expression plasmid pET28a possessing ORF of kshA2N This study
    pET28a-kshBN Expression plasmid pET28a possessing ORF of kshBN This study
    pMV261 Shuttle vector of Mycobacterium and E. coli carrying the heat shock (hsp60) promoter, Kanr W. R. Jacobs, Jr.
    p261-16srRNA pMV261 carrying the complete ORF of the 16S rRNA gene, for calibration in absolute quantification This study
    p261-kshA1N pMV261 derived, possessing kshA1N under the control of the hsp60 promoter for augmentation This study
    p261-kshA1V207T pMV261 derived, possessing kshA1V207T under the control of the hsp60 promoter for augmentation This study
    p261-kshA2N pMV261 derived, harboring Phsp60-kshA2N operon This study
    p261-kstD1 pMV261 derived, harboring Phsp60-kstD1 operon This study
Strains
    M. neoaurum
        ATCC 25795 Type strain, designated Mn25795 in this study ATCC
        MutkshA1N Mutant of Mn25795 with inactivation of KshA1N This study
        MutkshA2N Mutant of Mn25795 with inactivation of KshA2N This study
        Strain I MutkshA1NA2N, mutant of Mn25795 with inactivation of KshA1N and KshA2N This study
        MutkstDN Mutant of Mn25795 with inactivation of KstD1, KstD2, and KstD3 This study
    E. coli
        BL21-kshA1N&kshA2N Recombinant BL21(DE3) possessing pET28a-kshA1N or pET28a-kshA2N This study
        BL21-kshBN Recombinant BL21(DE3) possessing pET28a-kshBN This study
        BL21-kshA1V207T Recombinant BL21(DE3) possessing pET28a-kshA1V207T This study
        BL21-kshA1P210E Recombinant BL21(DE3) possessing pET28a-kshA1P210E This study
        BL21-kshA1D225R Recombinant BL21(DE3) possessing pET28a-kshA1D225R This study
        BL21-kshA1N238D Recombinant BL21(DE3) possessing pET28a-kshA1N238D This study
    Strain II Augmented strain containing p261kstD1, derived from strain I This study
    Strain III Augmented strain containing p261kstD1-hsd4A, derived from strain I This study
    Strain IV Augmented strain containing p261kshA1N, derived from MutkstDN Yao et al. (4)
    Strain V Augmented strain containing p261kshA1V207T, derived from MutkstDN This study
a

Mn25795 is a soil isolate deposited in the American Type Culture Collection.

For the conversion of sterol, two basic processes were induced, activation and proliferation of strains and implementation of sterol conversion. For the activation and proliferation of strains, colonies that formed on LB agar medium were usually inoculated into 5 ml of MYC/01 medium and then cultured up to an optical density at 600 nm (OD600) of 0.8 to 1.0; these cultures were directly used as seed cultures or further expanded to prepare 500 ml of seed culture for scale-up. For the implementation of sterol conversion, two methods were employed: growing cell transformation and resting cell transformation (26). Growing cell transformation was carried out in MYC/02 medium with the addition of cholesterol or phytosterols as described before (1, 25); the medium was inoculated with 10% (vol/vol) seed culture and grown at 30°C with shaking at 220 rpm or in a 7-liter bioreactor (7BG; Baoxing, Shanghai) at 250 rpm and airflow at 1 volume of air per unit of medium per minute (vvm). Resting cell transformation was carried out in 0.02 mM KH2PO4 buffer containing 20 g/liter phytosterols, 80 g/liter hydroxypropyl-β-cyclodextrin (HP-β-CD; RSC Chemical Industries Co. Ltd., Jiangsu, China), as previously described (26), and 100 g/liter wet mycobacterial cells harvested from the activation and proliferation culture of strains. Resting cell transformation was also performed under conditions similar to those used for growing cell transformation.

To determine the transcription and expression levels of kshAN, as well as the growth phenotypes of kshAN mutants, minimum medium (MM) [1.52 g liter−1 KH2PO4, 2.44 g liter−1 Na2HPO4, 0.5 g liter−1 (NH4)2SO4, 0.2 g liter−1 MgSO4 · 7H2O, 0.05 g liter−1 CaCl2 · 2H2O, pH 6.9] plus 18 mM glycerol (MM-glycerol) or 1.8 mM cholesterol (MM-sterol) was used. Culture in MM was also carried out at 30°C with shaking at 220 rpm, and growth was denoted by OD600 values that were determined against blank medium. To analyze the transcription of kshAN paralogous genes in Mn25795, cholesterol, rather than phytosterols, was used because it can be stably emulsified by Tween 80 in MM, which was diluted to 1.8 mM using stock solution (240 mM) emulsified by Tween 80 as described previously (1). To eliminate differences of Tween 80 between MM-glycerol medium and MM-sterol medium, 0.03% Tween 80 was also added into MM-glycerol medium. In addition, the induction of kshAN by AD was determined using 1.8 mM AD diluted from an 80 mM stock solution dissolved in dimethyl sulfoxide (DMSO). To eliminate the difference, an equal amount of DMSO was also added to MM-glycerol medium as a control. To assess the growth of mutants on sterol, 35 mg of cholesterol powder was directly added into MM with or without 18 mM glycerol as a control to exclude the putative toxicity of metabolites.

DNA and RNA extraction and transcriptional analysis of kshA homologues.

The isolation of genomic DNA from Mn25795 was conducted strictly according to Uhía et al. (27). The transcription levels of kshA homologues in wild-type, disrupted, and augmented strains were determined by reverse-transcription quantitative PCR (RT-qPCR). RNA extraction and real-time quantitative PCR analyses of cDNA samples were performed as previously reported (4). The quantity and quality of DNA and RNA were determined using a NanoDrop spectrophotometer (NanoDrop Technologies) and 2% (wt/vol) agarose gel electrophoresis. The specific RNA extraction process is given in the supplemental material. When the transcription of kshAN was analyzed in the engineered strains, the amount of kshAN mRNA present during cholesterol transformation was quantified in comparison with that of the parent strain ATCC 25795.

The following reverse transcription process was used: 2 μl of pd(N)6 random hexamers (10 μM), 1 μg of total RNA, and 2 μl of deoxynucleoside triphosphates (dNTPs) (10 mM) were first adjusted to 15 μl using double-distilled water (ddH2O), incubated at 70°C for 5 min, and immediately cooled on ice for 2 min. Twenty units of RNA inhibitor, 4 μl of 5× buffer with dithiothreitol (DTT), and 200 U of Super M-MLV reverse transcriptase then were added to 20 μl and incubated under the following conditions: 25°C, 10 min; 42°C, 60 min; and 95°C, 5 min. The mixture of cDNA was diluted to a final volume of 50 μl before quantification on a NanoDrop. Real-time qPCR analyses of cDNA samples were performed on a StepOne real-time PCR system (Applied Biosystems, CA). A 50-ng cDNA sample was completely mixed with 12.5 μl of SuperReal PreMix Plus with SYBR green I, 0.3 μM specific oligonucleotides (16SrRNA-f&r, RTqA1-f&r, and RTqA2-f&r) (Table 5), and 2 μl of ROX reference dye (for calibration of fluorescent signals from background and different wells) into 25 μl of reaction mixture. Amplification was performed with 15 min of predenaturing and then 40 cycles of 95°C for 10 s, 60°C for 30 s, and 72°C for 30 s. Melting curve analyses were performed at the end of each cycle. Each gene was measured in triplicate from three independent tests. The cDNA amplification efficiencies of samples, internal standards (16S rRNA), and calibrators (samples without cholesterol induction) were equivalently modulated, such that the relative amount of mRNA could be processed using the 2−△△CT algorithm (28). Each gene was determined in triplicate in three independent tests.

Bioinformatic analyses.

Comprehensive protein information on KshA is available at the UniProtKB database (http://www.uniprot.org/uniprot/P71875), and the mapping of its secondary structure was based on data provided from PDBsum services at the EMBL-EBI website (http://www.ebi.ac.uk/pdbsum/2ZYL).

The tertiary structures of two KshAs from Mn25795 (designated KshA1N and KshA2N, respectively) were modeled based on alignments of their coding sequences with that of KshAH37Rv, whose three-dimensional (3D) structure has been characterized and reported with the accession number 2ZYL in the PDB (http://www.ebi.ac.uk/pdbe-srv/view/entry/2zyl/summary.html) (3, 19). The SCWRL server assisted the analysis of differences between KshAN homologues and KshAH37Rv. These differences were visualized and analyzed using the PyMOL Molecular Graphics System.

Heterologous expression of Ksh enzymes in E. coli, protein purification, and activity assay.

In this study, the KshA and KshB components from Mn25795 (designated KshAN and KshBN, respectively) were expressed independently with a His tag at their C termini. E. coli cells harboring recombinant plasmids were developed as previously reported (4), and the primers used are shown in Table 5. Reconstituted cells were grown at 37°C at 220 rpm to an OD600 of 0.4 to 0.6 (for KshA1N and KshA2N) or OD600 of 0.2 (for an optimal soluble fraction of KshBN), followed by a 12-h induction using 0.5 mM isopropyl-β-d-thiogalactopyranoside (IPTG) at 30°C (for KshA1N) or 25°C (for KshA2N). The cell pellets were first independently harvested by centrifugation at 5,000 × g for 15 min at 4°C, resuspended in 10 ml of buffer A (50 mM Tris-HCl buffer, pH 7.5, 0.5 M NaCl, 20 mM imidazole) with 1 mM phenylmethylsulfonyl fluoride (PMSF), and disrupted using a French press under 20,000 lb/in2 pressure at 4°C (APV-2000-1; Germany). Final supernatants free of cell debris were obtained by centrifugation at 12,000 × g for 30 min at 4°C. As concluded by Petrusma et al. (19), the activity of KshA could not be maintained in vitro without KshB. Thus, the extracts of KshAN and KshBN and their premixed extracts were prepared for purification. Subsequent nickel-nitrilotriacetic acid (Ni-NTA) resin purification and dialysis steps were conducted at 4°C (29).

The Bradford method was used to determine the protein concentrations of purified KshAN and KshBN, with bovine serum albumin as the standard. Quantity One 1-D analysis software (version 4.62; Bio-Rad) was applied to determine the ratio of KshAN to KshBN in an SDS-PAGE image. Trace tracking in association with Gauss model bands, a standard processing method, assisted in eliminating background disturbances from identical lanes. Therefore, the absolute concentrations of KshAN and KshBN were quantified according to the signal ratio calculated by Quantity One.

Standard Ksh enzyme activity assay.

The KshBN assay system contained 0.25 mM NADH, 0.1 mM 2,6-dichlorophenolindophenol (DCPIP; J&K), and 1 to 2 μg of KshBN in 50 mM Tris-HCl (1 ml, pH 7.0). The system remained at room temperature for 15 min before thorough mixing. Assays were conducted by measuring the OD600 of DCPIP (ε = 21 mM−1 cm−1) at 22°C (5).

According to previous reports, the optimal Ksh activity assay mixture comprised 105 μM NADH, 200 μM steroid substrate dissolved in 100% isopropanol, and 50 mM Tris-HCl (200 μl, pH 7.0) (5, 11). Up to 10 μg of KshAN with an additional amount of KshBN to reach saturation (ultimately, the ratio of KshAN to KshBN was equal to 1:2) was included in the assay mixture. Thus, the specific activity of Ksh was characterized as the amount of steroids (equal to the oxidation of NADH) hydroxylated per minute using 1 mg of total Ksh protein at pH 7.0 and 35°C (μmol min−1 mg−1). The relationship between the substrate and reaction velocity was processed using SigmaPlot software (version 12.1) based on the Michaelis-Menten formula. The substrate concentration varied from 10 μM to 200 μM, and Vmax and Km were calculated according to the Lineweaver-Burk plot. Recording began upon the addition of substrate, which had been subjected to a 5-min preincubation at 35°C. A SpectraMax 190 (Molecular Devices) with the Soft-max PRO program was used to detect and analyze the oxidation of NADH at 340 nm (ε = 6.22 mM−1 cm−1) at 35°C (5, 12). The steroids dissolved poorly in water and were uniformly adjusted to 200 μM (25 μM for 4-cholestene-3-one) in the final mixture. The control was the same mixture without addition of substrate. Products with the 9α-hydroxyl group were visualized via high-performance liquid chromatography-ultraviolet (HPLC-UV) analysis.

Construction of chimeras of KshA homologues and mutants with amino acid substitutions.

The process of exchanging β-sheets and loop regions between KshA1N and KshA2N is shown in the supplemental material.

Site-directed mutagenesis of KshA1N.

The mutagenesis of KshA1N was developed with reference to the protein alignment of KshA1 and KshA5 from Rhodococcus rhodochrous DSM 43269 (19). A QuikChange site-directed mutagenesis kit (Stratagene) was employed to generate the kshA1V207T (GTC-ACC), kshA1P210E (CCG-GAG), kshA1D225R (GAT-CGT), and kshA1N238D (AAC-GAC) mutant strains.

Deletion of kshA1N and kshA2N in Mn25795 and complementation thereof.

Gene deletions in Mn25795 were performed as described in previous reports (1, 14). The process of developing suicide plasmids for genetic deletion and plasmids for functional complementation is shown in the supplemental material. Pairs of primers, kshA1N-del-U-f/kshA1N-del-D-r and kshA2N-del-U-f/kshA2N-del-D-r (Table S1), were employed for PCR amplification and sequenced for an expected knockout. Since no overlapping regions were found in the homologous arms of kshA1N and kshA2N, MutkshA1NA2N was achieved by the disruption of kshA2N on MutkshA1N.

In this study, complementation in mycobacteria was achieved by using pMV306-derived plasmids (p306kshA1N [ComkshA1N] and p306kshA2N), and a blank pMV306 plasmid was used as the control (Com-blank). Additionally, MutkshA1NA2N was complemented using p306kshA1N and p306kshA2N, creating ComkshA1&2N. All complemented strains with plasmids inserted were verified using PCR with the universal primers o-p306-f and o-p306-r. Complementation was confirmed by RT-qPCR to ensure the expression of complemented genes.

All plasmids constructed for gene knockout, overexpression, and complementation were electrointroduced into mycobacterial cells using electroporation protocols described previously (30).

Construction of augmented strains capable of producing 9-OHAD.

PCR products amplified from the plasmid pET28-kshA1V207T (Table 5) using the primers p261kshA1N-f and p261kshA1N-r were digested with DpnI (Fermentas) to remove the template, purified, further digested with BamHI-HindIII, and cloned into pMV261 to generate pMV261-kshA1V207T (Table 5). The primers and the plasmids used to construct other mutants are shown in Table 5. MutkstDN (inactivation of kstD1, kstD2, and kstD3 in Mn25795) and M. neoaurum NwIB-yV (pMV261-kshA1N transformed into MutkstDN, named strain IV in this study) were introduced according to Yao et al. (4). pMV261-kshA1V207T then was transformed into MutkstDN to create strain V. The primers o-p261-f and o-p261-r were used to confirm the introduction of those recombinant plasmids, and their expression levels were assessed by RT-qPCR.

Analyses of steroidal metabolites. (i) Sample preparation.

Phytosterol conversion in flasks either by growing cells or resting cells was sampled every 24 h. The growing cell transformation of phytosterols in the bioreactor was sampled every 12 h. The 9α-hydroxylation of various steroidal substrates by the reconstituted Ksh in vitro was sampled once the reaction was over. One milliliter of sample was extracted three times on a vortex mixer with 0.5 ml of ethyl acetate and 10% H2SO4 for 2 min before centrifugation at 12,000 × g for 20 min. At this point, thin-layer chromatography (TLC) was employed to initially evaluate the products, and gas chromatography (GC) was used to determine the levels of cholesterol or components of phytosterol mixture, as described previously (1). For liquid chromatography-mass spectrometry (LC-MS) analysis, the organic phase was transferred into a clean tube, dried under vacuum, and redissolved in 200 μl of methanol.

(ii) LC-MS.

9-OHAD dissolved in methanol was filtered through a 0.2-μm membrane and then analyzed by HPLC under the following chromatographic conditions: an Agilent XDB-C18 column (4.6 by 250 mm) with a constant temperature of 40°C, a mobile phase of methanol-water (70:30), and a flow rate of 1 ml min−1. The absorbance of UV254 represented the absolute concentration of products against the standard curve of 9-OHAD, ranging from 20 μM to 5 mM. The products of MutkshA1N transformation culture and the in vitro assay mixtures were analyzed by electrospray ionization-LC-mass spectrometry (LC-MS; Agilent 1100LC/MSD; Agilent Technologies) with DHT, ADD, AD, T, 6-OHAD, and 9-OHAD as authentic references. The MS parameters were the following: ion spray voltage, 30 V; ion source temperature, 100°C; desolvation temperature, 300°C; and a full scan from m/z 100 to m/z 1,000 in positive ionization mode.

GC-MS plus semipreparative HPLC.

The following analysis conditions were used for gas chromatography (GC)-MS: an Agilent ZORBAX Eclipse XDB C18 preparation column (4.6 by 150 mm; particle size, 5 μm) at a 1-ml min−1 flow rate with the methanol-water (7/3, vol/vol) mobile phase. Peaks were collected individually and evaporated under nitrogen flow. The pellets were redissolved in acetonitrile, derivatized at 60°C with an equivalent amount of BSTFA-1% TMCS for 30 min, and then analyzed on an Agilent 6890 series GC according to Yao et al. (4). An HP 5973 mass-selective detector was operated in standby in selected ion monitoring (SIM) mode with full scans from 100 m/z to 1,000 m/z. Molecular mass was determined against the trimethylsilyl derivatization reagents.

Accession number(s).

The genome sequencing information of Mn25795 has been deposited in the GenBank database with the accession number NZ_JMDW00000000. The accession numbers of kshA1, kshA2, kstD1, kstD2, and kstD3 gene sequences from Mn25795 are KF573736, KF573737, GQ411074.1, KF772209.1 and KF772210.1, respectively.

Supplementary Material

Supplemental material

ACKNOWLEDGMENT

This research was financially supported by the National Natural Science Foundation of China (no. 31370080).

Footnotes

Supplemental material for this article may be found at https://doi.org/10.1128/AEM.02777-17.

REFERENCES

  • 1.Yao K, Wang FQ, Zhang HC, Wei DZ. 2013. Identification and engineering of cholesterol oxidases involved in the initial step of sterols catabolism in Mycobacterium neoaurum. Metab Eng 15:75–87. doi: 10.1016/j.ymben.2012.10.005. [DOI] [PubMed] [Google Scholar]
  • 2.Malaviya A, Gomes J. 2008. Androstenedione production by biotransformation of phytosterols. Bioresour Technol 99:6725–6737. doi: 10.1016/j.biortech.2008.01.039. [DOI] [PubMed] [Google Scholar]
  • 3.Capyk JK, D'Angelo I, Strynadka NC, Eltis LD. 2009. Characterization of 3-ketosteroid 9α-hydroxylase, a Rieske oxygenase in the cholesterol degradation pathway of Mycobacterium tuberculosis. J Biol Chem 284:9937–9946. doi: 10.1074/jbc.M900719200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Yao K, Xu LQ, Wang FQ, Wei DZ. 2014. Characterization and engineering of 3-ketosteroid-Δ1-dehydrogenase and 3-ketosteroid-9α-hydroxylase in Mycobacterium neoaurum ATCC 25795 to produce 9α-hydroxy-4-androstene-3, 17-dione through the catabolism of sterols. Metab Eng 24:181–191. doi: 10.1016/j.ymben.2014.05.005. [DOI] [PubMed] [Google Scholar]
  • 5.Petrusma M, Dijkhuizen L, van der Geize R. 2009. Rhodococcus rhodochrous DSM 43269 3-ketosteroid 9α-hydroxylase, a two-component iron-sulfur-containing monooxygenase with subtle steroid substrate specificity. Appl Environ Microbiol 75:5300–5307. doi: 10.1128/AEM.00066-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.van der Geize R, Hessels G, Van Gerwen R, van der Meijden P, Dijkhuizen L. 2002. Molecular and functional characterization of kshA and kshB, encoding two components of 3-ketosteroid 9alpha-hydroxylase, a class IA monooxygenase, in Rhodococcus erythropolis strain SQ1. Mol Microbiol 45:1007–1018. doi: 10.1046/j.1365-2958.2002.03069.x. [DOI] [PubMed] [Google Scholar]
  • 7.van der Geize R, Hessels G, Nienhuis-Kuiper M, Dijkhuizen L. 2008. Characterization of a second Rhodococcus erythropolis SQ1 3-ketosteroid 9α-hydroxylase activity comprising a terminal oxygenase homologue, KshA2, active with oxygenase-reductase component KshB. Appl Environ Microbiol 74:7197–7203. doi: 10.1128/AEM.00888-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.van der Geize R, Yam K, Heuser T, Wilbrink MH, Hara H, Anderton MC, Sim E, Dijkhuizen L, Davies JE, Mohn WW. 2007. A gene cluster encoding cholesterol catabolism in a soil actinomycete provides insight into Mycobacterium tuberculosis survival in macrophages. Proc Natl Acad Sci U S A 104:1947–1952. doi: 10.1073/pnas.0605728104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Pandey AK, Sassetti CM. 2008. Mycobacterial persistence requires the utilization of host cholesterol. Proc Natl Acad Sci U S A 105:4376–4380. doi: 10.1073/pnas.0711159105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Hu YM, van der Geize R, Besra GS, Gurcha SS, Liu A, Rohde M, Singh M, Coates A. 2010. 3-Ketosteroid 9α-hydroxylase is an essential factor in the pathogenesis of Mycobacterium tuberculosis. Mol Microbiol 75:107–121. doi: 10.1111/j.1365-2958.2009.06957.x. [DOI] [PubMed] [Google Scholar]
  • 11.Petrusma M, Hessels G, Dijkhuizen L, van der Geize R. 2011. Multiplicity of 3-ketosteroid-9α-hydroxylase enzymes in Rhodococcus rhodochrous DSM43269 for specific degradation of different classes of steroids. J Bacteriol 193:3931–3940. doi: 10.1128/JB.00274-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Wilbrink MH, Petrusma M, Dijkhuizen L, van der Geize R. 2011. FadD19 of Rhodococcus rhodochrous DSM43269, a steroid-coenzyme A ligase essential for degradation of C-24 branched sterol side chains. Appl Environ Microbiol 77:4455–4464. doi: 10.1128/AEM.00380-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Andor A, Jekkel A, Hopwood DA, Jeanplong F, Ilkőy É, Kónya A, Kurucz I, Ambrus G. 2006. Generation of useful insertionally blocked sterol degradation pathway mutants of fast-growing mycobacteria and cloning, characterization, and expression of the terminal oxygenase of the 3-ketosteroid 9α-hydroxylase in Mycobacterium smegmatis mc2155. Appl Environ Microbiol 72:6554–6559. doi: 10.1128/AEM.00941-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Wei W, Wang FQ, Fan SY, Wei DZ. 2010. Inactivation and augmentation of the primary 3-ketosteroid-Δ1-dehydrogenase in Mycobacterium neoaurum NwIB-01: biotransformation of soybean phytosterols to 4-androstene-3,17-dione or 1,4-androstadiene-3,17-dione. Appl Environ Microbiol 76:4578–4582. doi: 10.1128/AEM.00448-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Schmidt CL, Shaw L. 2001. A comprehensive phylogenetic analysis of Rieske and Rieske-type iron-sulfur proteins. J Bioenerg Biomembr 33:9–26. doi: 10.1023/A:1005616505962. [DOI] [PubMed] [Google Scholar]
  • 16.Jiang HY, Parales RE, Lynch NA, Gibson DT. 1996. Site-directed mutagenesis of conserved amino acids in the alpha subunit of toluene dioxygenase: potential mononuclear non-heme iron coordination sites. J Bacteriol 178:3133–3139. doi: 10.1128/jb.178.11.3133-3139.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Kendall SL, Burgess P, Balhana R, Withers M, ten Bokum A, Lott JS, Gao C, Uhia-Castro I, Stoker NG. 2010. Cholesterol utilization in mycobacteria is controlled by two TetR-type transcriptional regulators: kstR and kstR2. Microbiology 156:1362–1371. doi: 10.1099/mic.0.034538-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Bragin EY, Shtratnikova VY, Dovbnya D, Schelkunov M, Pekov YA, Malakho S, Egorova O, Ivashina T, Sokolov S, Ashapkin V. 2013. Comparative analysis of genes encoding key steroid core oxidation enzymes in fast-growing Mycobacterium spp. strains. J Steroid Biochem Mol Biol 138:41–53. doi: 10.1016/j.jsbmb.2013.02.016. [DOI] [PubMed] [Google Scholar]
  • 19.Petrusma M, Dijkhuizen L, van der Geize R. 2012. Structural features in the KshA terminal oxygenase protein that determine substrate preference of 3-ketosteroid 9α-hydroxylase enzymes. J Bacteriol 194:115–121. doi: 10.1128/JB.05838-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Długoński J, Wilmańska D. 1998. Deleterious effects of androstenedione on growth and cell morphology of Schizosaccharomyces pombe. Antonie Van Leeuwenhoek 73:189–194. doi: 10.1023/A:1000640909646. [DOI] [PubMed] [Google Scholar]
  • 21.Capyk JK, Casabon I, Gruninger R, Strynadka NC, Eltis LD. 2011. Activity of 3-ketosteroid 9α-hydroxylase (KshAB) indicates cholesterol side chain and ring degradation occur simultaneously in Mycobacterium tuberculosis. J Biol Chem 286:40717–40724. doi: 10.1074/jbc.M111.289975. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Petrusma M, van der Geize R, Dijkhuizen L. 2014. 3-Ketosteroid 9α-hydroxylase enzymes: Rieske non-heme monooxygenases essential for bacterial steroid degradation. Antonie Van Leeuwenhoek 106:157–172. doi: 10.1007/s10482-014-0188-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Xu LQ, Liu YJ, Yao K, Liu HH, Tao XY, Wang FQ, Wei DZ. 2016. Unraveling and engineering the production of 23, 24-bisnorcholenic steroids in sterol metabolism. Sci Rep 6:21928. doi: 10.1038/srep21928. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Shtratnikova VY, Schelkunov MI, Dovbnya DV, Pekov YA, Bragin EY, Ashapkin VV, Donova MV. 2015. Complete genome sequence of Mycobacterium sp. strain VKM Ac-1817D, capable of producing 9α-hydroxy-androst-4-ene-3, 17-dione from phytosterol. Genome Announc 3:e01447-14. doi: 10.1128/genomeA.01447-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Xiong LB, Liu HH, Xu LQ, Wei DZ, Wang FQ. 2017. Role identification and application of SigD in the transformation of soybean phytosterol to 9α-hydroxy-4-androstene-3,17-dione in Mycobacterium neoaurum. J Agric Food Chem 65:626–631. doi: 10.1021/acs.jafc.6b05314. [DOI] [PubMed] [Google Scholar]
  • 26.Gao XQ, Feng JX, Hua Q, Wei DZ, Wang XD. 2014. Investigation of factors affecting biotransformation of phytosterols to 9-hydroxyandrost-4-ene-3,-17-dione based on the HP-β-CD-resting cells reaction system. Biocatal Biotransformation 32:343–347. doi: 10.3109/10242422.2014.976633. [DOI] [Google Scholar]
  • 27.Uhía I, Galán B, Morales V, García J. 2011. Initial step in the catabolism of cholesterol by Mycobacterium smegmatis mc2155. Environ Microbiol 13:943–959. doi: 10.1111/j.1462-2920.2010.02398.x. [DOI] [PubMed] [Google Scholar]
  • 28.Livak KJ, Schmittgen TD. 2001. Analysis of relative gene expression data using real-time quantitative PCR and the 2−ΔΔCT method. Methods 25:402–408. doi: 10.1006/meth.2001.1262. [DOI] [PubMed] [Google Scholar]
  • 29.Li B, Wang W, Wang FQ, Wei DZ. 2010. Cholesterol oxidase ChoL is a critical enzyme that catalyzes the conversion of diosgenin to 4-ene-3-keto steroids in Streptomyces virginiae IBL-14. Appl Microbiol Biotechnol 85:1831–1838. doi: 10.1007/s00253-009-2188-0. [DOI] [PubMed] [Google Scholar]
  • 30.Gordhan BG, Parish T. 2001. Gene replacement using pretreated DNA. Methods Mol Med 54:77–92. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplemental material

Articles from Applied and Environmental Microbiology are provided here courtesy of American Society for Microbiology (ASM)

RESOURCES