Next Article in Journal
Isolation and Characterization of Low-Temperature and High-Salinity Amylase from Halomonas sp. KS41843
Previous Article in Journal
Exploring Exopolysaccharides Produced in Indigenous Mexican Fermented Beverages and Their Biotechnological Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of 3,4-Dihydroxybenzoic Acid in E. coli and C. glutamicum Using Dehydroshikimate Dehydratase of Different Types

Ajinomoto-Genetika Research Institute, 117545 Moscow, Russia
*
Author to whom correspondence should be addressed.
Fermentation 2025, 11(8), 464; https://doi.org/10.3390/fermentation11080464
Submission received: 5 June 2025 / Revised: 25 July 2025 / Accepted: 5 August 2025 / Published: 12 August 2025
(This article belongs to the Section Microbial Metabolism, Physiology & Genetics)

Abstract

Dehydroshikimate (DHS) dehydratase (DSD) catalyzes the conversion of DHS into 3,4-dihydroxybenzoic acid (3,4-DHBA), a compound with promising applications across various industries. The DSD from Podospora anserina (DSDPa) was characterized and its catalytic properties were compared with those of previously investigated enzymes, AsbF (Bacillus thuringiensis), Qa-4 (Neurospora crassa), and QsuB (Corynebacterium glutamicum), both in vitro and in vivo using tube fermentation. Escherichia coli and C. glutamicum were used as platforms to construct model 3,4-DHBA producers. To increase DHS availability in both hosts, shikimate dehydrogenase AroE was inactivated, and the plasmid pVS7-aroG4, encoding 3-deoxy-D-arabinoheptulosonate 7-phosphate synthase (E. coli), was introduced. In E. coli, heterologous 3,4-DHBA synthesis was achieved through chromosomal integration of dsd genes. The fungal genes were codon-optimized for this bacterium. The same genes were cloned into the pVK9 vector and introduced into C. glutamicum, where 3,4-DHBA degradation was disrupted (ΔpcaHG). AsbF (kcat ~ 1 s−1) showed poor 3,4-DHBA accumulation in both hosts (1–1.5 g/L). The enzymes with better catalytic characteristics, QsuB (kcat ~ 60 s−1), DSDPa (kcat ~ 125 s−1), and Qa-4 (kcat ~ 220 s−1), provided 5 g/L 3,4-DHBA in E. coli and 3 g/L 3,4-DHBA in C. glutamicum, except for Qa-4. The low production (~1.5 g/L) observed for Qa-4 in C. glutamicum might be attributed to a non-optimal nucleotide sequence rich in codons rare for C. glutamicum.

1. Introduction

3,4-Dihydroxybenzoic acid (3,4-DHBA), also known as protocatechuic acid, is a natural metabolite. In soil microorganisms, 3,4-DHBA serves as an intermediate in quinate/shikimate catabolism. 3,4-DHBA is a phenolic acid with functional –COOH and –OH groups, which can form complexes with divalent and trivalent cations. Consequently, in some Bacillus spp., 3,4-DHBA acts as an iron-chelating component of the siderophore [1,2,3].
In recent decades, 3,4-DHBA has been found to possess antioxidant, neuroprotective, and anti-inflammatory properties [4,5]. This makes it possible to use 3,4-DHBA in cosmetic and pharmaceutical industries [6]. Moreover, 3,4-DHBA was shown to be used as an intermediate for the synthesis of other valuable compounds such as vanillin, catechol, adipic acid, muconic acid, and biopolymers [7,8,9,10,11,12]. Current methods for 3,4-DHBA production have several disadvantages. Extraction from plants gives a low yield and, subsequently, high cost of the target compound [13,14]. Chemical synthesis implies an energy-consuming process and environmentally harmful reagents [15]. Therefore, microbial fermentation of renewable raw materials can be applied for green and sustainable 3,4-DHBA production.
The construction of 3,4-DHBA-producing strains has been described previously [16,17,18,19,20,21,22]. Such strains were created based on C. glutamicum, Escherichia coli, Bacillus licheniformis, and Pseudomonas putida. The 3,4-DHBA synthesis from D-xylose using C. glutamicum as a host has been described [21]. However, other 3,4-DHBA-producing strains utilized glucose as a carbon source.
Two biosynthetic routes for microbial 3,4-DHBA production have been established: dehydration of 3-dehydroshikimate (DHS), an intermediate of the common aromatic pathway, or a two-step conversion of chorismate (CHA), the end product of this pathway (Figure 1). The latter route may be less favorable due to additional requirements for ATP, PEP, and NADPH. Interestingly, 3,4-DHBA synthesis via CHA appeared to be effective when combined with DHS-based synthesis in a C. glutamicum producing strain. This approach achieved the highest reported titer to date (~ 82 g/L 3,4-DHBA) under aerobic growth-arrested fermentation conditions [22].
Here, only 3,4-DHBA synthesis by DHS dehydratase (EC: 4.2.1.118) (DSD) was considered, as we aimed to compare DSDs from different sources. Based on their structural properties, DHS dehydratases have been divided into five types: bacterial one-domain, bacterial two-domain, fungal one-domain, bacterial membrane-associated, and QuiC2-like enzymes [23,24]. Previously, we compared 3,4-DHBA production in vitro and in vivo using enzymes of the first three types and employing E. coli as a platform strain [17,25]. Two-domain QsuB from C. glutamicum and one-domain Qa-4 from the filamentous fungus Neurospora crassa showed higher specific activity at physiological pH 7.5 and provided better 3,4-DHBA accumulation than the one-domain AsbF from Bacillus thuringiensis. In this work, we biochemically characterized DSD from Podospora pauciseta/anserina (DSDPa) for the first time. DSDPa possessed 72% identity with Qa-4 and demonstrated catalytic properties similar to those of Qa-4 and QsuB.
C. glutamicum has been described as a host for 3,4-DHBA synthesis, primarily using its native enzyme QsuB [21,22]. In this work, we compared DHS dehydratases of different types in a C. glutamicum platform strain. Unlike E. coli, C. glutamicum both synthesizes and degrades 3,4-DHBA as an intermediate of the quinate/shikimate utilization pathway (Figure 1B). Therefore, 3,4-DHBA degradation was disrupted in the C. glutamicum model producer. To reduce DHS flux into the common aromatic pathways, the aroE gene, encoding the main biosynthetic shikimate dehydrogenase, was inactivated in both host strains. The investigated bacteria possess additional minor enzymes with this catalytic function (Figure 1). YdiB exhibits low activity in E. coli, and its inactivation does not affect the growth in M9 minimal medium [26]. Quinate/shikimate dehydrogenase QsuD is essential for utilizing these compounds as carbon sources in C. glutamicum [27]. To enhance the carbon flux into the aromatic pathways and subsequent precursor DHS accumulation, we introduced a plasmid encoding E. coli 3-deoxy-D-arabinoheptulosonate 7-phosphate (DAHP) synthase (pVS7-aroG4). Thus, we aimed to reveal more promising enzymes for 3,4-DHBA production.

2. Materials and Methods

2.1. Strains, Plasmids and Growth Conditions

E. coli and C. glutamicum strains and plasmids are listed in Table 1. E. coli and C. glutamicum cells were cultivated with aeration (240 rpm) in LB (Tryptone, 10; NaCl, 10; Yeast extract, 5 (g/L)) and BHI (Thermo Scientific™, Detroit, MI, USA) at 37 °C and 30 °C, respectively. To prepare plates, agar (20 g/L) was added. Antibiotics were used in the following concentrations (mg/L): ampicillin (Ap), 200, tetracycline (Tc), 12.5, spectinomycin (Spe), 50 (E. coli); chloramphenicol (Cm), 6, kanamycin (Km), 25, spectinomycin (Spe), 50, gentamicin (Gm), 1.5 (C. glutamicum).
C. glutamicum cells were grown (~2 h) in BHI supplemented with glycine (2%) and Tween 80 (0.1%) before electroporation.

2.2. Fermentation

Fermentation of E. coli and C. glutamicum strains was performed in tubes (18 mm × 200 mm), containing 2 mL of fermentation medium. Tubes were inoculated with 0.2 mL of seed culture. To prepare seed culture, one loop (∅ 3 mm) of cells taken from a fresh Petri dish was introduced into tubes (13 mm × 150 mm) containing 3 mL of medium and incubated as described below.
The E. coli fermentation medium contained (g/L) glucose, 40; CaCO3, 60; tryptone, 10; NaCl, 10; yeast extract, 5; (NH4)2SO4, 0.5; KH2PO4, 0.5; MgSO4∙7H2O, 0.3; FeSO4∙7H2O, 0.005; thiamine, 0.01; and 4-hydroxybenzoic acid, para-aminobenzoic acid, and 2,3-dihydroxybenzoic acid, 0.005 (pH 7.0). Fermentation was performed at 34 °C and 240 rpm for 44 h. The seed culture was grown in LB at 34 °C with aeration (240 rpm) for 3 h.
The C. glutamicum fermentation medium contained (g/L): glucose, 50; MgSO4·7H2O, 0.4; CaCO3, 60; (NH4)2SO4, 5; KH2PO4, 2; FeSO4∙7H2O, 0.01; MnSO4∙4H2O, 0.01; ZnSO4∙7H2O, 0.0072; soybean hydrolysate, 0.75 (Ajinomoto Co., Inc., Kawasaki, Japan); biotin, 0.00012; thiamine, 0.01; para-aminobenzoic acid, 0.005; L-Phe, 1; L-Tyr, 1; L-Trp, 1. Fermentation was performed at 30 °C and 240 rpm until complete glucose consumption. The seed medium (pH 7.5) contained (g/L) glucose, 5; tryptone, 10; yeast extract, 10; KH2PO4, 1; MgSO4·7H2O, 0.4; FeSO4·7H2O, 0.01; MnSO4·5H2O, 0.01; urea, 3; and biotin 0.0001. The seed culture was grown for 18 h at 30 °C, 240 rpm.
Residual glucose was tested by GlukoPHAN test strips (Erba Lachema, Brno, Czech Republic) and concentrations were determined with BIOSEN C_line (EKF-diagnostic GmbH, Barleben, Germany). To make samples transparent for optical density (OD) measurements, 2 N HCl was added to convert CaCO3 present in the fermentation medium into CaCl2. DHS and 3,4-DHBA concentrations were analyzed by HPLC as described previously [17].

2.3. Strains Construction

The integration of the dsdPa gene into the E. coli chromosome was performed as described previously for genes asbF, qa-4, and qsuB [17]. For this purpose, the integrative vector pAH162-λattL-TcRattR-PlacUV5-dsdPa was created. Helper plasmids pAH123 and pMW-int-xis were used for integration into the native φ80attB site of MG1655ΔaroE chromosome and for subsequent elimination of the Tc resistance marker, respectively (Figure 2A).
Gene deletions in the chromosome of C. glutamicum were performed using RecET-mediated recombineering [31]. Helper plasmids pVC-AprR-lacI-Ptrc-id2-recE564T and p06-Pdap-cre were used for replacement of the target gene by the marker gene and marker elimination, respectively (Figure 2B).

2.4. DNA Manupulations

All DNA procedures, including PCR and Circular Polymerase Extension Cloning (CPEC), were performed as described previously [17,25]. The DNA sequence of dsdPa was optimized for E. coli based on amino acid sequence CAP65279.1 and chemically synthesized (Twist Bioscience, South San Francisco, CA, USA). The obtained gene was used for expression in both E. coli and C. glutamicum. Primers for PCR used in this study are listed in Table S1.
DNA fragments used for the gene replacement (aroE (cgl1629), pcaHG, and qsuB) were obtained by overlap PCR of three fragments. These included the excisable marker lox71-cat-lox66 and two flanking homologous arms (~800 bp) targeting chromosomal regions upstream and downstream the insertion site. Plasmid pMW119-lox66-cat-lox71 served as the template for marker amplification.
All plasmids were obtained by CPEC [35] using pre-amplified DNA fragments with overlapping ends. The DNA of chemically synthesized asbF, qa-4, and dsdPa genes was used as templates. Chromosomal DNA from AJ1511 [28] and VKPM B-10747 [36] strains served for amplifying the qsuB and aroG4 genes, respectively. To obtain the vector parts, DNA of pVK9-lacI-Ptrc-id2-gfp, pAH162-λattL-TcRattR-PlacUV5-qsuB, pET22b, and pVS7 plasmids was used as the templates after digestion with PstI, PvuII, SalI, and BamHI, respectively. Translation initiation regions (included in primers) for all genes inserted into the pVK9 vector were designed to ensure similar expression levels using UTR Designer [37]. The expression indices of dsd genes were similar in both E. coli and C. glutamicum, as determined by GenScript (https://www.genscript.com/tools/rare-codon-analysis, accessed on 17 January 2025). The resulting plasmids were verified by sequencing (Evrogen, Moscow, Russia).

2.5. Enzyme Activity Assays

The activity was measured as previously described [25,38] by monitoring the change in absorbance of 3,4-DHBA at 290 nm (ε290 = 3.89 × 103 M−1 × sm−1) for 0.5 min using Genesys10S UV-visible spectrophotometer (Thermo Scientific, Madison, WI, USA). The standard reaction was performed at 25 °C in a quartz cuvette in 1 mL of reaction mixture with the following composition: 0.1 M Tris/HCl buffer (pH 7.5), 10 mM metal salt (CoCl2, MgCl2, and MnCl2), 0.1–5 mM DHS and purified enzyme (150 nM AsbF, 10 nM Qa-4, 20 nM QsuB, 45 nM DSDPa).
To exclude the potential influence of the His-tag on enzyme activity, DHS dehydratase activity was analyzed in crude extracts of BL21(DE3)/pET22b-dsdPa and BL21(DE3)/pET22b-dsdPa-His6 strains (Figure S1). Crude protein extracts were prepared using xTractor-Buffer (Takara Bio, San Jose, CA, USA). Since the activities were identical (mean statistical error < 0.5%), the purified C-terminally His6-tagged DSDPa protein was used for subsequent kinetic analyses. The His6-tagged protein was purified using the Capturem™ His-Tagged Purification Miniprep Kit (Takara Bio, San Jose, CA, USA). To remove residual divalent cations, the enzyme was incubated with 1 mM EDTA on ice for 1 h prior to activity assays. For metal cofactor testing, 1 mM DHS was used. The Km and Vmax values of DSDPa were determined by non-linear regression analysis [39]. The inhibition type was investigated by the “quotient velocity plot” method [40]. For this analysis, DHS was used at concentrations 0.2, 0.4, and 1 mM, while 3,4-DHBA was evaluated as a potential inhibitor at concentrations ranging from 0 to 0.6 mM.

2.6. Sequence Alignment

The amino acid sequences of DHS dehydratases, QuiC from Acinetobacter baylyi (AAC37159.1), QuiC1 from P. putida (AAN68163.1), QuiC2 from Listeria monocytogenes (CAD00312.1), DSDPa (CAP65279.1), Qa-4 (CAA32750.1), QsuB (BAB97815.1), and AsbF (AOM10649.1), were obtained from the protein database (https://www.ncbi.nlm.nih.gov/protein, accessed on 15 April 2024). Analysis of homologous sequences was performed using the blastp algorithm (https://www.uniprot.org/blast, accessed on 15 April 2024), and phylogenetic tree rendering (https://www.uniprot.org/align, accessed on 15 April 2024) was conducted using UniProt services.

2.7. Statistical Analysis

All values in tables represent arithmetic means of at least three independent experiments, with errors given as confidence interval. Microsoft Excel 2010 was used for calculations. To assess the significance of differences between the investigated enzymes, we performed Analysis of Variance (ANOVA) followed by Tukey’s HSD (honestly significant difference) test. This statistical analysis was applied to both catalytic properties (Table S2) and 3,4-DHBA accumulation in test tube fermentations (Tables S3–S5).

3. Results

3.1. DHS Dehydratase from P. anserina and Its Expression in T7 System

The DHS dehydratase from P. anserina (DSDPa) and Qa-4 both belong to type IV DHS dehydratases (Figure 3). DSDPa was applied for vanillin production in Schizosacharomyces pombe, though it was not characterized biochemically [7]. The precursor of vanillin producer with the dsdPa gene accumulated ~ 0.3 g/L 3,4-DHBA, while E. coli producer expressing the qa-4 gene produced ~ 3 g/L 3,4-DHBA [17]. These results were obtained in different host organisms in separate studies and did not provide an insight into the diversity of enzymes of the same type.
To enable direct comparison between DSDPa, Qa-4, and the enzymes of other types, we characterized DSDPa using the same experimental approach previously applied to QsuB, AsbF, and Qa-4 [17,25].

3.2. Catalytic Properties of DSDPa and Their Comparison with Those of AsbF, Qa-4, QsuB

DSD is a metal-dependent enzyme, and we measured DSDPa activity in the presence of various divalent metal ions (Mg2+, Ca2+, Zn2+, Co2+ and Mn2+). Unlike previously tested enzymes [25], DSDPa retained its activity after incubation with EDTA, which was used to remove residual metal ions from the active site (Figure 4A). DSDPa was slightly inhibited in the presence of Ca2+, Zn2+ and Mg2+, while Mn2+ and Co2+, especially, activated the enzyme. Therefore, Co2+ was a preferable cofactor. All further experiments were conducted with 10 mM CoCl2.
The optimal pH for DSDPa was in the range of 8.1–8.5, similar to QsuB (pH 8.0–8.4) (Figure 4B). Since we intended to use the enzyme in bacterial cells, we studied the enzyme activities at physiological pH 7.5.
The kinetic parameters of DSDPa (Table 2) were derived from the corresponding kinetic curve (Figure 5).
DSDPa had higher substrate specificity, intermediate meaning of kcat and the best Keff, when compared to other catabolic enzymes, QsuB and Qa-4 (Table 2). According to ANOVA and Tukey’s HSD tests, Keff of DSDPa and Qa-4 were statistically similar, while that of QsuB was closer to the less active AsbF (Table S2). kcat and Keff characterize the enzyme activity at high and low substrate concentrations, respectively. The fungal enzymes DSDPa and Qa-4, being phylogenetically related, exhibited the most favorable kinetic properties, suggesting they would be particularly effective for 3,4-DHBA production.
The enzyme’s susceptibility to inhibition by the target compound is a crucial parameter for biotechnological applications. Previous studies [17] demonstrated that AsbF showed the highest sensitivity to 3,4-DHBA inhibition compared to QsuB and Qa-4. Our analysis revealed that DSDPa undergoes mixed-type inhibition by 3,4-DHBA (Figure 6), like it was determined previously for QsuB (Figure 6). 3,4-DHBA could bind to both the free enzyme (Ki) and the enzyme-substrate complex (K’i). The inhibition constants were similar in magnitude to those reported for QsuB (Table 3).

3.3. 3,4-DHBA Production in E. coli Strains

The E. coli MG1655ΔaroE strain naturally accumulated DHS. In our previous work [17], we have already used derivatives of this strain with integrated DHS dehydratase genes to compare 3,4-DHBA production.
The isogenic MG1655∆aroE PlacUV5-dsdPa strain was additionally constructed in this work. All strains were evaluated in test tube fermentation (Table 4).
All investigated genes were under the control of the PlacUV5 promoter, which normally requires IPTG induction for transcription. However, strains expressing the more active enzymes, Qa-4, QsuB, and DSDPa, accumulated 3,4-DHBA even without IPTG (Table 4, lines 3–5), likely due to promoter leakage. Notably, IPTG addition did not significantly increase 3,4-DHBA accumulation (Table 4, lines 8–10). As previously reported, the MG1655∆aroE PlacUV5-asbF strain showed limited 3,4-DHBA production, resulting from both low enzyme activity and strong product inhibition by 3,4-DHBA [17]. To increase DHS supply, we introduced plasmid pVS7-aroG4, which carries the mutant variant of E. coli 3-deoxy-D-arabinoheptulosonate 7-phosphate (DAHP) synthase that is resistant to L-Phe feedback inhibition [36]. Strain with the asbF gene and the parent MG1655ΔaroE accumulated significant amounts of DHS (~ 7 g/L) after introduction of pVS7-aroG4. The plasmid strains expressing Qa-4, QsuB, and DSDPa accumulated about 5 g/L 3,4-DHBA. These results demonstrate that Qa-4, QsuB, and DSDPa show comparable performance in the engineered E. coli platform strain MG1655ΔaroE/pVS7-aroG4 (Table S3).

3.4. 3,4-DHBA Production in C. glutamicum Strains

To assess DHS dehydratase activity in C. glutamicum, we constructed plasmids pVK9-lacI-Ptrc-id2-dsd (Figure S2), designed to maintain similar expression levels of dsd genes. These plasmids, along with the empty pVK9 vector, were transformed into the K118 strain. The resulting strains were evaluated in test tube fermentation (Table 5).
The K118 strain harboring the empty pVK9 vector accumulated ~ 0.7 g/L 3,4-DHBA due to the expression of the native chromosomal qsuB gene (Table 5, lines 1, 6, 11). Strains carrying plasmids with qsuB and dsdPa genes produced 3,4-DHBA even without IPTG addition, likely due to promoter leakage and the high activity of these enzymes (Table 5, lines 4, 5). This conclusion was supported by the prolonged cultivation time (CT), extended from 24 to 29 hours until complete glucose consumption, and by reduced biomass accumulation (OD) under IPTG addition (Table 5, lines 9, 10). Strains expressing AsbF and Qa-4 demonstrated lower 3,4-DHBA accumulation, exceeding the basal production of the K118 strain only in the presence of IPTG (Table 5, lines 6–8).
Following the approach used for E. coli 3,4-DHBA producers, C. glutamicum strains were transformed with pVS7-aroG4. Both 3,4-DHBA accumulation and biomass increased in strains expressing QsuB and DSDPa; however, DHS was still absent (Table 5, lines 14–15). This suggests that the substrate DHS might be limited in these strains. Strains expressing AsbF and Qa-4 did not show an increase in 3,4-DHBA titer and continued to accumulate DHS (Table 5, lines 12–13). Therefore, Qa-4 activity was as low as that of AsbF in C. glutamicum (Table S4). In the K118 strain, 3,4-DHBA production was also supported by native QsuB. To assess whether the native enzyme might influence the activity of heterologous ones, we analyzed 3,4-DHBA production in a ΔqsuB derivative of the K118 strain.

3.5. 3,4-DHBA Production in C. glutamicumΔqsuB Strains

Inactivation of the native qsuB gene reduced 3,4-DHBA accumulation in all plasmid strains containing dsd genes (Table 6, lines 2–5; 7–10). The K118ΔqsuB expressing QsuB and DSDPa showed decreased biomass accumulation under IPTG induction, and introduction of the pVS7-aroG4 plasmid did not restore biomass levels (Table 6, lines 14–15). These results suggest that DHS availability was reduced in the ΔqsuB strain.
Deletion of the qsuB gene might have affected the translation of the downstream aroD (qsuC) gene, which is located in the same operon (Figure 7). Although the deletion was designed to keep the reading frame, the removed DNA region may have contained additional transcription signals, or the inserted DNA fragment may have attenuated aroD transcription.
The strains expressing Qa-4 and AsbF accumulated higher levels of DHS than the strains expressing QsuB and DSDPa (Table 6 and Table S5). Therefore, the activities of Qa-4 and AsbF were lower than those of QsuB and DSDPa in the ΔqsuB strain.

3.6. Sequence Comparison of qa-4 and dsdPa Genes

The observed difference in Qa-4 and DSDPa activities in E. coli and C. glutamicum may stem from variations in gene expression efficiency between these hosts. The presence of codons with a usage frequency less than 10 in C. glutamicum was analyzed in qa-4 and dsdPa genes (Table 7). The usage frequencies of some of these codons differed by two times or more in C. glutamicum and E. coli (indicated in bold in Table 7). In a whole, the qa-4 gene sequence contained more codons problematic for C. glutamicum than dsdPa. The most noticeable difference was observed for Leu codon UUA (27 in qa-4 versus 17 in dsdPa) rare for C. glutamicum. Thus, poor expression of Qa-4 in C. glutamicum may be considered as a major factor underlying its reduced enzymatic activity and consequent low 3,4-DHBA production.

4. Discussion

Relatively few studies have reported comparative screening of DHS dehydratases [11,18]. The natural diversity of these enzymes enables selection of optimal candidates for microbiological synthesis, depending on specific hosts requirements. In previous work, we demonstrated that Qa-4 and QsuB enzymes, related to quinate/shikimate catabolic pathway, exhibited higher specific activity and greater resistance to 3,4-DHBA inhibition compared to AsbF, a siderophore biosynthesis enzyme [17]. In this study, we characterized another catabolic DHS dehydratase from P. anserina. Both DSDPa and Qa-4, which originate from closely related fungi, showed superior catalytic efficiency, kcat and Keff values, compared to QsuB (Table 2). Surprisingly, no significant differences were observed between the E. coli MG1655ΔaroE-based strains expressing these enzymes, even when DHS supply was enhanced through introduction of the pVS7-aroG4 plasmid (see Figure 8 for final comparison of the enzymes and host strains). All strains expressing Qa-4, DSDPa and QsuB reached 3,4-DHBA accumulation up to 5 g/L. Presumably, the differences and in vitro kinetic advantages of a particular enzyme may become apparent in optimized high-performance strains or under fed-batch fermentation conditions.
Significant functional differences emerged between phylogenetically related fungal enzymes when expressed in C. glutamicum. The expression of the qa-4 gene in C. glutamicum appears problematic and may explain the observed 3,4-DHBA production impairment. Both qa-4 and dsdPa genes were chemically synthesized with codon optimization for E. coli expression. While E. coli and C. glutamicum genes are typically used interchangeably without optimization, sequence analysis revealed the qa-4 gene contains substantially more codons that are rare in C. glutamicum compared to dsdPa. The presence of rare codons in native sequences is known to modulate translational kinetics, facilitating proper protein folding [41]. However, when heterologous rare codons are incompatible with the host’s tRNA pool, this can result in translational stalling, premature termination, or protein misfolding. Notably, in contrast to E. coli, DSDPa demonstrated marginally better performance than QsuB in the presence of pVS7-aroG4 plasmid in C. glutamicum platform strain.
Parallel experiments with 3,4-DHBA-producing strains revealed the differences in metabolic networks in E. coli and C. glutamicum. C. glutamicum platform strain had a bottleneck in DHS accumulation, even in the presence of the pVS7-aroG4 plasmid. E. coli MG1655ΔaroE/pVS7-aroG4 accumulated DHS up to 7 g/L (Table 4, line 11), while C. glutamicum K118/ pVS7-aroG4 produced less than 1 g/L DHS (Table 5, line 11). The observed DHS accumulation deficiency in C. glutamicum platform strain may result from metabolic flux diversion through downstream reactions of the common aromatic pathway. The minor shikimate dehydrogenase QsuD is more active than nearly silenced YdiB of E. coli. At the same time, the qsuC gene, which encodes the DHS-synthesizing dehydroquinate dehydratase (see Figure 1), is part of the qsu operon. This operon is known to be induced by shikimate or chorismate in C. glutamicum [27,42], explaining its suboptimal expression in our K118 strain. The DHS deficiency and consequent decrease in 3,4-DHBA production were further exacerbated in the ΔqsuB strain, likely due to the possible polar effect of this modification on qsuC transcription (Figure 8). In contrast, aroD expression is constitutive in the E. coli MG1655 strain [43]. Thus, C. glutamicum platform strain requires metabolic engineering to achieve 3,4-DHBA production levels comparable to E. coli.
Based on our findings, DSDPa is suitable for heterologous 3,4-DHBA biosynthetic pathways creation in both E. coli and C. glutamicum.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/fermentation11080464/s1, Figure S1: SDS-PAGE of crude cell extracts and purified DSDPa; Figure S2: The scheme of pVK9-lacI-Ptrc-id2-dsd; Table S1: Primers used in the work. Table S2. Statistical evaluation of difference between catalytic parameters of DHS dehydratases using initial data from Table 3. Table S3. Statistical evaluation of 3,4-DHBA production by the MG1655ΔaroE-based strains using initial data from Table 4. Table S4. Statistical evaluation of 3,4-DHBA production by the K118-based strains using initial data from Table 5. Table S5. Statistical evaluation of 3,4-DHBA production by the K118ΔqsuB-based strains using initial data from Table 6.

Author Contributions

Conceptualization, E.S. and V.D.; Methodology, E.S.; Software, E.S. and A.K.; Validation, E.S. and A.K.; Formal analysis, A.K. and N.N.; Investigation, A.K., N.N. and E.S.; Resources, E.S. and V.D.; Data curation, E.S. and V.D.; Writing—original draft preparation, A.K. and E.S.; Writing—review and editing, V.D. and N.S.; Visualization, A.K.; Supervision, E.S. and V.D.; Project administration, V.D. and N.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The analyzed data presented in this study are included within this article. Further data are available on reasonable request from the corresponding author.

Acknowledgments

The authors would like to thank Elizaveta Fedorova for HPLC analysis.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Wilson, M.K.; Abergel, R.J.; Raymond, K.N.; Arceneaux, J.E.; Byers, B.R. Siderophores of Bacillus anthracis, Bacillus cereus, and Bacillus thuringiensis. Biochem. Biophys. Res. Commun. 2006, 348, 320–325. [Google Scholar] [CrossRef]
  2. Fox, D.T.; Hotta, K.; Kim, C.Y.; Koppisch, A.T. The missing link in petrobactin biosynthesis: asbF encodes a (−)-3-dehydroshikimate dehydratase. Biochemistry 2008, 47, 12251–12253. [Google Scholar] [CrossRef] [PubMed]
  3. Pfleger, B.F.; Kim, Y.; Nusca, T.D.; Maltseva, N.; Lee, J.Y.; Rath, C.M.; Scaglione, J.B.; Janes, B.K.; Anderson, E.C.; Bergman, N.H.; et al. Structural and functional analysis of AsbF: Origin of the stealth 3, 4-dihydroxybenzoic acid subunit for petrobactin biosynthesis. Proc. Natl. Acad. Sci. USA 2008, 105, 17133–17138. [Google Scholar] [CrossRef] [PubMed]
  4. Kakkar, S.; Bais, S. A review on protocatechuic acid and its pharmacological potential. Int. Sch. Res. Not. 2014, 2014, 952943. [Google Scholar] [CrossRef] [PubMed]
  5. Krzysztoforska, K.; Mirowska-Guzel, D.; Widy-Tyszkiewicz, E. Pharmacological effects of protocatechuic acid and its therapeutic potential in neurodegenerative diseases: Review on the basis of in vitro and in vivo studies in rodents and humans. Nutr. Neurosci. 2019, 22, 72–82. [Google Scholar] [CrossRef]
  6. Lin, C.Y.; Huang, C.S.; Huang, C.Y.; Yin, M.C. Anticoagulatory, antiinflammatory, and antioxidative effects of protocatechuic acid in diabetic mice. J. Agric. Food Chem. 2009, 57, 6661–6667. [Google Scholar] [CrossRef]
  7. Hansen, E.H.; Møller, B.L.; Kock, G.R.; Bünner, C.M.; Kristensen, C.; Jensen, O.R.; Okkels, F.T.; Olsen, C.E.; Motawia, M.S.; Hansen, J. De novo biosynthesis of vanillin in fission yeast (Schizosaccharomyces pombe) and baker’s yeast (Saccharomyces cerevisiae). Appl. Environ. Microbiol. 2009, 75, 2765–2774. [Google Scholar] [CrossRef]
  8. Draths, K.M.; Frost, J.W. Environmentally compatible synthesis of catechol from D-glucose. J. Am. Chem. Soc. 1995, 117, 2395–2400. [Google Scholar] [CrossRef]
  9. Pugh, S.; McKenna, R.; Osman, M.; Thompson, B.; Nielsen, D.R. Rational engineering of a novel pathway for producing the aromatic compounds p-hydroxybenzoate, protocatechuate, and catechol in Escherichia coli. Process Biochem. 2014, 49, 1843–1850. [Google Scholar] [CrossRef]
  10. Niu, W.; Draths, K.M.; Frost, J.W. Benzene-free synthesis of adipic acid. Biotechnol. Prog. 2002, 18, 201–211. [Google Scholar] [CrossRef]
  11. Weber, C.; Brückner, C.; Weinreb, S.; Lehr, C.; Essl, C.; Boles, E. Biosynthesis of cis, cis-muconic acid and its aromatic precursors, catechol and protocatechuic acid, from renewable feedstocks by Saccharomyces cerevisiae. Appl. Environ. Microbiol. 2012, 78, 8421–8430. [Google Scholar] [CrossRef]
  12. Fang, X.; Guo, X.; Tang, W.; Gu, Q.; Wu, Y.; Sun, H.; Gao, J. Efficient toughening of DGEBA with a bio-based protocatechuic acid derivative. ACS Omega 2023, 8, 9962–9968. [Google Scholar] [CrossRef] [PubMed]
  13. Yang, Y.C.; Wei, M.C.; Huang, T.C.; Lee, S.Z. Extraction of protocatechuic acid from Scutellaria barbata D. Don using supercritical carbon dioxide. J. Supercrit. Fluid. 2013, 81, 55–66. [Google Scholar] [CrossRef]
  14. Antony, F.M.; Wasewar, K. Reactive extraction: A promising approach to separate protocatechuic acid, Environ. Sci. Pollut. R. 2019, 27, 27345–27357. [Google Scholar] [CrossRef] [PubMed]
  15. Pearl, I.A. Reactions of vanillin and its derived compounds. I. The reaction of vanillin with silver oxide1. J. Am. Chem. Soc. 1946, 68, 429–432. [Google Scholar] [CrossRef]
  16. Okai, N.; Miyoshi, T.; Takeshima, Y.; Kuwahara, H.; Ogino, C.; Kondo, A. Production of protocatechuic acid by Corynebacterium glutamicum expressing chorismate-pyruvate lyase from Escherichia coli. Appl. Microbiol. Biotechnol. 2016, 100, 135–145. [Google Scholar] [CrossRef]
  17. Shmonova, E.A.; Savrasova, E.A.; Fedorova, E.N.; Doroshenko, V.G. Comparative Analysis of Catabolic and Anabolic Dehydroshikimate Dehydratases for 3,4-DHBA Production in Escherichia coli. Microorganisms 2022, 10, 1357. [Google Scholar] [CrossRef]
  18. Rao, Y.; Wang, J.; Zhao, R.; Zhan, Y.; Ma, X.; He, P.; Cai, D.; Chen, S. Efficient sustainable production of protocatechuic acid from glucose by engineered Bacillus licheniformis. Chem. Eng. J. 2025, 505, 159320. [Google Scholar] [CrossRef]
  19. Li, J.; Fu, J.; Yue, C.; Shang, Y.; Ye, B.C. Highly efficient biosynthesis of protocatechuic acid via recombinant Pseudomonas putida KT2440. J. Agric. Food Chem. 2023, 71, 10375–10382. [Google Scholar] [CrossRef]
  20. Örn, O.E.; Sacchetto, S.; van Niel, E.W.; Hatti-Kaul, R. Enhanced protocatechuic acid production from glucose using Pseudomonas putida 3-dehydroshikimate dehydratase expressed in a phenylalanine-overproducing mutant of Escherichia coli. Front. Bioeng. Biotechnol. 2021, 9, 695704. [Google Scholar] [CrossRef]
  21. Labib, M.; Görtz, J.; Brüsseler, C.; Kallscheuer, N.; Gätgens, J.; Jupke, A.; Marienhagen, J.; Noack, S. Metabolic and process engineering for microbial production of protocatechuate with Corynebacterium glutamicum. Biotechnol. Bioeng. 2021, 118, 4414–4427. [Google Scholar] [CrossRef]
  22. Kogure, T.; Suda, M.; Hiraga, K.; Inui, M. Protocatechuate overproduction by Corynebacterium glutamicum via simultaneous engineering of native and heterologous biosynthetic pathways. Metab. Eng. 2021, 65, 232–242. [Google Scholar] [CrossRef]
  23. Peek, J.; Roman, J.; Moran, G.R.; Christendat, D. Structurally diverse dehydroshikimate dehydratase variants participate in microbial quinate catabolism. Mol. Microbiol. 2017, 103, 39–54. [Google Scholar] [CrossRef] [PubMed]
  24. Xue, K.; Prezioso, S.M.; Christendat, D. QuiC2 represents a functionally distinct class of dehydroshikimate dehydratases identified in Listeria species including Listeria monocytogenes. Environ. Microbiol. 2020, 22, 2680–2692. [Google Scholar] [CrossRef] [PubMed]
  25. Shmonova, E.A.; Voloshina, O.V.; Ovsienko, M.V.; Smirnov, S.V.; Nolde, D.E.; Doroshenko, V.G. Characterization of the Corynebacterium glutamicum dehydroshikimate dehydratase QsuB and its potential for microbial production of protocatechuic acid. PLoS ONE 2020, 15, e0231560. [Google Scholar] [CrossRef] [PubMed]
  26. Michel, G.; Roszak, A.W.; Sauvé, V.; Maclean, J.; Matte, A.; Coggins, J.R.; Cygler, M.; Lapthorn, A.J. Structures of shikimate dehydrogenase AroE and its paralog YdiB: A common structural framework for different activities. J. Biol. Chem. 2003, 278, 19463–19472. [Google Scholar] [CrossRef]
  27. Kubota, T.; Tanaka, Y.; Takemoto, N.; Watanabe, A.; Hiraga, K.; Inui, M.; Yukawa, H. Chorismate-dependent transcriptional regulation of quinate/shikimate utilization genes by LysR-type transcriptional regulator QsuR in Corynebacterium glutamicum: Carbon flow control at metabolic branch point. Mol. Microbiol. 2014, 92, 356–368. [Google Scholar] [CrossRef]
  28. Nishio, Y.; Koseki, C.; Tonouchi, N.; Matsui, K.; Sugimoto, S.; Usuda, Y. Analysis of strain-specific genes in glutamic acid-producing Corynebacterium glutamicum ssp. lactofermentum AJ 1511. J. Gen. Appl. Microbiol. 2017, 63, 157–164. [Google Scholar] [CrossRef]
  29. Minaeva, N.I.; Gak, E.R.; Zimenkov, D.V.; Skorokhodova, A.Y.; Biryukova, I.V.; Mashko, S.V. Dual-In/Out strategy for genes integration into bacterial chromosome: A novel approach to step-by-step construction of plasmid-less marker-less recombinant E. coli strains with predesigned genome structure. BMC Biotechnol. 2008, 8, 63. [Google Scholar] [CrossRef]
  30. Katashkina, Z.; Skorokhodova, A.; Zimenkov, D.V.; Gulevich, A.; Minaeva, N.I.; Doroshenko, V.G.; Biriukova, I.V.; Mashko, S.V. Tuning of expression level of the genes of interest located in the bacterial chromosome. Mol. Biol. 2005, 39, 823–831. [Google Scholar] [CrossRef]
  31. Lobanova, J.S.; Gorshkova, N.V.; Krylov, A.A.; Stoynova, N.V.; Mashko, S.V. Genome engineering of the Corynebacterium glutamicum chromosome by the Extended Dual-In/Out strategy. J. Microbiol. Methods 2022, 200, 106555. [Google Scholar] [CrossRef]
  32. Gorshkova, N.V.; Lobanova, J.S.; Tokmakova, I.L.; Smirnov, S.V.; Akhverdyan, V.Z.; Krylov, A.A.; Mashko, S.V. Mu-driven transposition of recombinant mini-Mu unit DNA in the Corynebacterium glutamicum chromosome. Appl. Microbiol. Biotechnol. 2018, 102, 2867–2884. [Google Scholar] [CrossRef] [PubMed]
  33. Nakamura, J.S.; Hirano, S.; Ito, H. L-glutamic Acid Producing Microorganism and a Method for Producing L-glutamic Acid. U.S. Patent 7,794,989, 14 September 2010. [Google Scholar]
  34. Nishio, Y.; Yamamoto, Y.; Yamada, K.; Yokota, K. Method for Producing Target Substance by Fermentation. U.S. Patent 9,045,789, 2 June 2015. [Google Scholar]
  35. Quan, J.; Tian, J. Circular polymerase extension cloning of complex gene libraries and pathways. PLoS ONE 2009, 4, e6441. [Google Scholar] [CrossRef] [PubMed]
  36. Doroshenko, V.G.; Tsyrenzhapova, I.S.; Krylov, A.A.; Kiseleva, E.M.; Ermishev, V.Y.; Kazakova, S.M.; Biryukova, I.V.; Mashko, S.V. Pho regulon promoter-mediated transcription of the key pathway gene aroG Fbr improves the performance of an l-phenylalanine-producing Escherichia coli strain. Appl. Microbiol. Biotechnol. 2010, 88, 1287–1295. [Google Scholar] [CrossRef] [PubMed]
  37. Seo, S.W.; Yang, J.S.; Kim, I.; Yang, J.; Min, B.E.; Kim, S.; Jung, G.Y. Predictive design of mRNA translation initiation region to control prokaryotic translation efficiency. Metab. Eng. 2013, 15, 67–74. [Google Scholar] [CrossRef]
  38. Strøman, P.; Reinert, W.R.; Giles, N.H. Purification and characterization of 3-dehydroshikimate dehydratase, an enzyme in the inducible quinic acid catabolic pathway of Neurospora crassa. J. Biol. Chem. 1978, 253, 4593–4598. [Google Scholar] [CrossRef]
  39. Ritchie, R.J.; Prvan, T. Current statistical methods for estimating the Km and Vmax of Michaelis-Menten kinetics. Biochem. Educ. 1996, 24, 196–206. [Google Scholar] [CrossRef]
  40. Yoshino, M.; Murakami, K. A graphical method for determining inhibition constants. J. Enzym. Inhib. Med. Chem. 2009, 24, 1288–1290. [Google Scholar] [CrossRef]
  41. Kim, S.J.; Yoon, J.S.; Shishido, H.; Yang, Z.; Rooney, L.A.; Barral, J.M.; Skach, W.R. Translational tuning optimizes nascent protein folding in cells. Science 2015, 348, 444–448. [Google Scholar] [CrossRef]
  42. Teramoto, H.; Inui, M.; Yukawa, H. Regulation of expression of genes involved in quinate and shikimate utilization in Corynebacterium glutamicum. Appl. Environ. Microbiol. 2009, 75, 3461–3468. [Google Scholar] [CrossRef]
  43. Moore, L.R.; Caspi, R.; Boyd, D.; Berkmen, M.; Mackie, A.; Paley, S.; Karp, P.D. Revisiting the y-ome of Escherichia coli. Nucleic Acids Res. 2024, 52, 12201–12207. [Google Scholar] [CrossRef]
Figure 1. 3,4-DHBA synthesis by DHS dehydratase (A) and aromatic pathways in E. coli (B) and C. glutamicum (C). Common aromatic pathway reactions are shown in black. Genes deleted in 3,4-DHBA-producing strains are shown in red. Native quinate/shikimate utilization pathway is shown in blue in (C). The reactions for 3,4-DHBA synthesis via CHA are shown in green for heterologous enzyme and in orange for native enzyme. Abbreviations of compounds: E4P—erythrose 4-phosphate; PEP—phosphoenolpyruvate; DAHP—3-deoxy-D-arabino-heptulosonate-7-phosphate; DHQ—3-dehydroquinate; QA—quinate; DHS—3-dehydroshikimate; SHIK—shikimate, CHA—chorismate; 4-HBA—4-hydroxybenzoate. Genes encoding enzymes of common aromatic pathway: aroF, aroG, aroH—DAHP-synthases; aroB—3-dehydroquinate synthase; aroD—dehydroquinate dehydratase; aroE—shikimate dehydrogenase. Genes encoding enzymes of quinate/shikimate utilization pathway: qsuA—quinate/shikimate transporter; qsuB—DHS dehydratase; qsuD—quinate/shikimate dehydrogenase; qsuC—3-dehydroquinate dehydratase; pcaHG—3,4-dioxygenase. Enzymes of 3,4-DHBA synthesis via CHA: UbiC—chorismate pyruvate lyase; PobA—4-hydroxybenzoate 3-monooxygenase.
Figure 1. 3,4-DHBA synthesis by DHS dehydratase (A) and aromatic pathways in E. coli (B) and C. glutamicum (C). Common aromatic pathway reactions are shown in black. Genes deleted in 3,4-DHBA-producing strains are shown in red. Native quinate/shikimate utilization pathway is shown in blue in (C). The reactions for 3,4-DHBA synthesis via CHA are shown in green for heterologous enzyme and in orange for native enzyme. Abbreviations of compounds: E4P—erythrose 4-phosphate; PEP—phosphoenolpyruvate; DAHP—3-deoxy-D-arabino-heptulosonate-7-phosphate; DHQ—3-dehydroquinate; QA—quinate; DHS—3-dehydroshikimate; SHIK—shikimate, CHA—chorismate; 4-HBA—4-hydroxybenzoate. Genes encoding enzymes of common aromatic pathway: aroF, aroG, aroH—DAHP-synthases; aroB—3-dehydroquinate synthase; aroD—dehydroquinate dehydratase; aroE—shikimate dehydrogenase. Genes encoding enzymes of quinate/shikimate utilization pathway: qsuA—quinate/shikimate transporter; qsuB—DHS dehydratase; qsuD—quinate/shikimate dehydrogenase; qsuC—3-dehydroquinate dehydratase; pcaHG—3,4-dioxygenase. Enzymes of 3,4-DHBA synthesis via CHA: UbiC—chorismate pyruvate lyase; PobA—4-hydroxybenzoate 3-monooxygenase.
Fermentation 11 00464 g001
Figure 2. Schemes for gene integration in E. coli (A) and gene deletion in C. glutamicum (B). Genome engineering of E. coli and C. glutamicum chromosomes was performed using genetic tools described previously [29,30,31,32]. For E. coli, the site-specific recombination systems of E. coli phages φ80 and λ were used for gene integration (helper plasmid pAH123) and marker elimination (pMW-int-xis), respectively. For C. glutamicum, the homologous recombination system of E. coli Rac prophage RecE564/RecT (C-terminus of RecE starting at residue 564) and E. coli P1 Cre-mediated site-specific recombination system were used for target gene replacement with the marker (helper plasmid pVC-AprR-lacI-Ptrc-id2-recE564T) and marker elimination, respectively (helper plasmid p06-PdapA-cre).
Figure 2. Schemes for gene integration in E. coli (A) and gene deletion in C. glutamicum (B). Genome engineering of E. coli and C. glutamicum chromosomes was performed using genetic tools described previously [29,30,31,32]. For E. coli, the site-specific recombination systems of E. coli phages φ80 and λ were used for gene integration (helper plasmid pAH123) and marker elimination (pMW-int-xis), respectively. For C. glutamicum, the homologous recombination system of E. coli Rac prophage RecE564/RecT (C-terminus of RecE starting at residue 564) and E. coli P1 Cre-mediated site-specific recombination system were used for target gene replacement with the marker (helper plasmid pVC-AprR-lacI-Ptrc-id2-recE564T) and marker elimination, respectively (helper plasmid p06-PdapA-cre).
Fermentation 11 00464 g002
Figure 3. A simplified phylogenetic tree for five distinct classes of DHS dehydratases. I—bacterial membrane-associated, II—bacterial one-domain, III—QuiC2-like, IV—fungal one-domain, and V—bacterial two-domain enzymes. The DHS dehydratases examined in this study are indicated. All DSD groups were supported with bootstrap values exceeding 0.95 (from 100 replicates).
Figure 3. A simplified phylogenetic tree for five distinct classes of DHS dehydratases. I—bacterial membrane-associated, II—bacterial one-domain, III—QuiC2-like, IV—fungal one-domain, and V—bacterial two-domain enzymes. The DHS dehydratases examined in this study are indicated. All DSD groups were supported with bootstrap values exceeding 0.95 (from 100 replicates).
Fermentation 11 00464 g003
Figure 4. Comparison of DSDPa and QsuB by specificity to metal cofactors (A) and by pH dependence (B).
Figure 4. Comparison of DSDPa and QsuB by specificity to metal cofactors (A) and by pH dependence (B).
Fermentation 11 00464 g004
Figure 5. The kinetic curve of DSDPa. Data fitting to Michaelis–Menten kinetics is shown in purple line. The green dashed line indicates Km value on the x-axis and half-maximal velocity at y-axis.
Figure 5. The kinetic curve of DSDPa. Data fitting to Michaelis–Menten kinetics is shown in purple line. The green dashed line indicates Km value on the x-axis and half-maximal velocity at y-axis.
Fermentation 11 00464 g005
Figure 6. Quotient velocity plots for the determination of DSDPa inhibition type. Experimental data were plotted as (Vmax–v)/v versus 3,4-DHBA concentration (0–0.6 mM) at three substrate (DHS) concentrations: 0.2 mM, 0.4 mM, and 1 mM (solid lines). Vmax and v correspond to maximal velocity and reaction velocity at given inhibitor concentration, respectively. Linear fitting (dashed lines) was used to determine the interception point of the plots.
Figure 6. Quotient velocity plots for the determination of DSDPa inhibition type. Experimental data were plotted as (Vmax–v)/v versus 3,4-DHBA concentration (0–0.6 mM) at three substrate (DHS) concentrations: 0.2 mM, 0.4 mM, and 1 mM (solid lines). Vmax and v correspond to maximal velocity and reaction velocity at given inhibitor concentration, respectively. Linear fitting (dashed lines) was used to determine the interception point of the plots.
Fermentation 11 00464 g006
Figure 7. qsu locus of C. glutamicum containing qsuB and aroD (qsuC).
Figure 7. qsu locus of C. glutamicum containing qsuB and aroD (qsuC).
Fermentation 11 00464 g007
Figure 8. Comparison of 3,4-DHBA accumulation in E. coli and C. glutamicum strains expressing DHS dehydratases of different types. The figure sums up the data from Table 4, Table 5 and Table 6.
Figure 8. Comparison of 3,4-DHBA accumulation in E. coli and C. glutamicum strains expressing DHS dehydratases of different types. The figure sums up the data from Table 4, Table 5 and Table 6.
Fermentation 11 00464 g008
Table 1. Bacterial strains and plasmids used in this work.
Table 1. Bacterial strains and plasmids used in this work.
Strain or PlasmidRelevant CharacteristicsSource
E. coli strains
BL21(DE3)F–ompT gal dcm lon hsdSB(rBmB) λ(DE3 [lacI lacUV5-T7p07 ind1 sam7 nin5]) [malB+]K−12S)Novagen (Merck Millipore, Darmstadt, Germany)
MG1655∆aroEF λ ilvG rfb-50 rph-1ΔaroE[25]
MG1655∆aroE PlacUV5-qsuBattφ803,4-DHBA producing strains[17]
MG1655∆aroE PlacUV5-asbFattφ80
MG1655∆aroE PlacUV5-qa-4attφ80
MG1655∆aroE PlacUV5-dsdPaattφ80This work
C. glutamicum strains
AJ1511ATCC13869 lacking cryptic plasmid pAM330[28]
K118AJ1511ΔaroEΔpcaHGThis work
K118ΔqsuBThis work
Plasmids
pAH162-λattL-TcRattR-
PlacUV5-qsuB
TcR, oriRγ, phi80attP; used for the creation of integrative vector with the dsdPa gene[17]
pAH162-λattL-TcRattR-PlacUV5-dsdPaTcR, oriRγ, phi80attP; used for PlacUV5-dsdPa integration into E. coli chromosomeThis work
pAH123ApR, the helper for the PlacUV5-dsdPa integration[29]
pMW-int-xisApR, oriR101, repA101ts, λcIts857, λPR → λxis-int; the helper for marker removal in E. coli.[30]
pMW119-lox66-cat-lox71CmR; contained excisable Cm-marker. lox71/lox66—mutant lox sites that exclude chromosomal deletions occurring after further introduction of identical markers. A. Krylov
(JSC “AGRI”)
pVC-AprR-lacI-Ptrc-id2-recE564TAprR, ori pMB1, ori pAM330; the helper for RecET-mediated recombination in C. glutamicum[31]
p06-PdapA-creGmR, ori pMB1, ori pCG1; the helper for Cre-mediated marker removal in C. glutamicum[32]
pET22b-dsdPa-His6ApR; production of DSDPa protein with and without His-tagThis work
pET22b-dsdPa
pVK9KmR; ori pCG1, ori ColE1[33]
pVK9-lacI-Ptrc-id2-gfpKmR; oripCG1, oriColE1, lacI, IPTG inducible Ptrc-id2A. Krylov (JSC “AGRI”)
pVK9-lacI-Ptrc-id2-asbFKmR; plasmids containing dsd genesThis work
pVK9-lacI-Ptrc-id2-qa-4
pVK9-lacI-Ptrc-id2-qsuB
pVK9-lacI-Ptrc-id2-dsdPa
pVS7SpeR; ori pMB1, repBI1[34]
pVS7-aroG4SpeR; ori pMB1, repBI1, aroG4This work
Table 2. Kinetic properties of DHS dehydratases.
Table 2. Kinetic properties of DHS dehydratases.
EnzymeKm, μMkcat, s−1Keff, 10−3/μM/sSource
DSDPa236 ± 29125 ± 7527 ± 167This work
QsuB960 ± 8061 ± 160 ± 12[25]
Qa-4600 ± 20219 ± 1370 ± 70[17]
AsbF36 ± 71.1 ± 0.129 ± 7
Table 3. Inhibition constants of DSDPa and QsuB.
Table 3. Inhibition constants of DSDPa and QsuB.
EnzymeKi, mMK’i, mMSource
DSDPa0.33 ± 0.040.61 ± 0.07This work
QsuB~0.38~0.96[17]
Table 4. Accumulation of 3,4-DHBA and DHS in the culture broth of the E. coli MG1655∆aroE-based strains containing different genes for DSD in test tube fermentation.
Table 4. Accumulation of 3,4-DHBA and DHS in the culture broth of the E. coli MG1655∆aroE-based strains containing different genes for DSD in test tube fermentation.
NDSDpVS7-aroG41 mM IPTGOD540DHS3,4-DHBARes. Glc *
g/L
138 ± 11.5 ± 0.1<0.13.0 ± 0.1
2AsbF33 ± 1<0.1<0.14.0 ± 0.1
3Qa-432 ± 1<0.11.1 ± 0.1
4QsuB33 ± 10.4 ± 0.10.7 ± 0.1
5DSDPa33 ± 1<0.10.9 ± 0.1
6+33 ± 11.3 ± 0.1<0.12.5 ± 0.1
7AsbF33 ± 1<0.10.2 ± 0.15.5 ± 0.1
8Qa-432 ± 1<0.11.1 ± 0.1
9QsuB31 ± 1<0.11.2 ± 0.1
10DSDPa32 ± 1<0.11.2 ± 0.1
11++37 ± 17.2 ± 0.1<0.1
12AsbF43 ± 17.3 ± 0.70.5 ± 0.1
13Qa-437 ± 1<0.14.9 ± 0.1
14QsuB38 ± 1<0.14.8 ± 0.1
15DSDPa38 ± 2<0.15.0 ± 0.5
* Res. Glc—residual glucose determined after cultivation time (CT) = 44 h.
Table 5. Accumulation of 3,4-DHBA and DHS in the culture broth of the K118/pVK9-lacI-Ptrc-id2-dsd strains in test tube fermentation.
Table 5. Accumulation of 3,4-DHBA and DHS in the culture broth of the K118/pVK9-lacI-Ptrc-id2-dsd strains in test tube fermentation.
NDSD *pVS7-aroG41 mM IPTGOD540DHS3,4-DHBACT **, h
g/L
185 ± 10.6 ± 0.10.7 ± 0.124
2AsbF84 ± 10.6 ± 0.10.7 ± 0.1
3Qa-481 ± 10.5 ± 0.10.6 ± 0.1
4QsuB90 ± 1<0.12.7 ± 0.1
5DSDPa88 ± 1<0.12.4 ± 0.1
6+85 ± 10.5 ± 0.10.7 ± 0.124
7AsbF84 ± 20.2 ± 0.11.5 ± 0.3
8Qa-483 ± 30.2 ± 0.11.2 ± 0.1
9QsuB62 ± 2<0.12.2 ± 0.229
10DSDPa72 ± 1<0.12.1 ± 0.1
11++87 ± 10.8 ± 0.10.7 ± 0.124
12AsbF83 ± 20.3 ± 0.11.3 ± 0.1
13Qa-486 ± 10.3 ± 0.11.4 ± 0.2
14QsuB82 ± 6<0.12.7 ± 0.127
15DSDPa86 ± 1<0.13.4 ± 0.124
* The control strain “−” contained pVK9 vector. K118/pVK9 and K118/pVK9/pVS7 strains had the similar performance. ** CT—cultivation time. Fermentation was terminated when glucose (40 g/L) was consumed.
Table 6. Accumulation of 3,4-DHBA and DHS in the culture broth of the C. glutamicum K118ΔqsuB/pVK9-lacI-Ptrc-id2-dsd strains in test tube fermentation.
Table 6. Accumulation of 3,4-DHBA and DHS in the culture broth of the C. glutamicum K118ΔqsuB/pVK9-lacI-Ptrc-id2-dsd strains in test tube fermentation.
NDSD *pVS7-aroG41 mM IPTGOD540DHS3,4-DHBARes. Glc **
g/L
170 ± 71.0 ± 0.1<0.10
2AsbF75 ± 10.9 ± 0.10.1 ± 0.1
3Qa-469 ± 10.9 ± 0.10.1 ± 0.1
4QsuB71 ± 1<0.10.6 ± 0.1
5DSDPa65 ± 10.8 ± 0.10.3 ± 0.1
6+68 ± 10.9 ± 0.10.1 ± 0.10
7AsbF71 ± 40.3 ± 0.10.4 ± 0.1
8Qa-470 ± 20.4 ± 0.10.6 ± 0.1
9QsuB38 ± 1<0.10.7 ± 0.11.8 ± 0.1
10DSDPa40 ± 3<0.10.6 ± 0.11.6 ± 0.2
11++67 ± 10.9 ± 0.10.1 ± 0.10
12AsbF71 ± 30.4 ± 0.10.6 ± 0.1
13Qa-472 ± 10.5 ± 0.10.6 ± 0.1
14QsuB33 ± 1<0.10.6 ± 0.11.6 ± 0.1
15DSDPa43 ± 1<0.10.8 ± 0.11.5 ± 0.1
* The control strain “−” contained pVK9 vector. K118ΔqsuB/pVK9 and K118ΔqsuB/pVK9/pVS7 strains had the similar performance. ** CT = 24 h for strains with Res. Glc = 0 and CT = 28 h for strains with Res. Glc.
Table 7. Comparative analysis of codon usage frequencies for qa-4 and dsdPa genes in E. coli and C. glutamicum.
Table 7. Comparative analysis of codon usage frequencies for qa-4 and dsdPa genes in E. coli and C. glutamicum.
AACodonC. glutamicum FrequencyE. coli FrequencyNumber of
Codons in qa-4
Number of
Codons in dsdPa
Arg (R)CGA6.64.33-
CGG4.94.11319
AGA2.21.425
AGG3.21.600
Cys (C)UGU2.35.955
UGC4.18.022
Gly (G)GGG6.78.632
His (H)CAU6.715.854
Ile (I)AUA1.83.731
Leu (L)UUA5.115.22717
CUA5.85.352
Pro (P)CCC9.76.430
Val (V)GUA8.111.501
SerAGU4.97.200
UCA8.27.800
UCG7.68.071
TyrUAU7.416.864
ThrACA7.66.410
ACG8.81.500
∑ codons rare for C. glutamicum relatively E. coli4833
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Shmonova, E.; Kruglova, A.; Nikandrov, N.; Stoynova, N.; Doroshenko, V. Synthesis of 3,4-Dihydroxybenzoic Acid in E. coli and C. glutamicum Using Dehydroshikimate Dehydratase of Different Types. Fermentation 2025, 11, 464. https://doi.org/10.3390/fermentation11080464

AMA Style

Shmonova E, Kruglova A, Nikandrov N, Stoynova N, Doroshenko V. Synthesis of 3,4-Dihydroxybenzoic Acid in E. coli and C. glutamicum Using Dehydroshikimate Dehydratase of Different Types. Fermentation. 2025; 11(8):464. https://doi.org/10.3390/fermentation11080464

Chicago/Turabian Style

Shmonova, Ekaterina, Arina Kruglova, Nikita Nikandrov, Nataliya Stoynova, and Vera Doroshenko. 2025. "Synthesis of 3,4-Dihydroxybenzoic Acid in E. coli and C. glutamicum Using Dehydroshikimate Dehydratase of Different Types" Fermentation 11, no. 8: 464. https://doi.org/10.3390/fermentation11080464

APA Style

Shmonova, E., Kruglova, A., Nikandrov, N., Stoynova, N., & Doroshenko, V. (2025). Synthesis of 3,4-Dihydroxybenzoic Acid in E. coli and C. glutamicum Using Dehydroshikimate Dehydratase of Different Types. Fermentation, 11(8), 464. https://doi.org/10.3390/fermentation11080464

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop