Skip to main content

Production of p-amino-l-phenylalanine (l-PAPA) from glycerol by metabolic grafting of Escherichia coli

Abstract

Background

The non-proteinogenic aromatic amino acid, p-amino-l-phenylalanine (l-PAPA) is a high-value product with a broad field of applications. In nature, l-PAPA occurs as an intermediate of the chloramphenicol biosynthesis pathway in Streptomyces venezuelae. Here we demonstrate that the model organism Escherichia coli can be transformed with metabolic grafting approaches to result in an improved l-PAPA producing strain.

Results

Escherichia coli K-12 cells were genetically engineered for the production of l-PAPA from glycerol as main carbon source. To do so, genes for a 4-amino-4-deoxychorismate synthase (pabAB from Corynebacterium glutamicum), and genes encoding a 4-amino-4-deoxychorismate mutase and a 4-amino-4-deoxyprephenate dehydrogenase (papB and papC, both from Streptomyces venezuelae) were cloned and expressed in E. coli W3110 (lab strain LJ110). In shake flask cultures with minimal medium this led to the formation of ca. 43 ± 2 mg l−1 of l-PAPA from 5 g l−1 glycerol. By expression of additional chromosomal copies of the tktA and glpX genes, and of plasmid-borne aroFBL genes in a tyrR deletion strain, an improved l-PAPA producer was obtained which gave a titer of 5.47 ± 0.4 g l−1 l-PAPA from 33.3 g l−1 glycerol (0.16 g l-PAPA/g of glycerol) in fed-batch cultivation (shake flasks). Finally, in a fed-batch fermenter cultivation, a titer of 16.7 g l−1 l-PAPA was obtained which is the highest so far reported value for this non-proteinogenic amino acid.

Conclusion

Here we show that E. coli is a suitable chassis strain for l-PAPA production. Modifying the flux to the product and improved supply of precursor, by additional gene copies of glpX, tkt and aroFBL together with the deletion of the tyrR gene, increased the yield and titer.

Background

The general aromatic biosynthesis pathway (also called shikimate pathway) in most microorganisms and plants starts by condensation of the precursors phosphoenolpyruvate (PEP) and erythrose-4-phosphate (E4P) to provide the first intermediate, 3-deoxy-d-arabino-heptulosonate-7-phosphate (DAHP). Eventually (after introduction of another PEP molecule), the last common metabolite, chorismate [1] is formed. From chorismate, various biosynthetic pathways diverge which give rise to a plethora of aromatic compounds. Among them are three proteinogenic aromatic amino acids [l-phenylalanine (l-Phe), l-tyrosine (l-Tyr), and l-tryptophan (l-Trp)] and aromatic vitamins such as quinones, folate (via p-aminobenzoate), and p-hydroxybenzoate [2,3,4,5]. Quinones (ubiquinone, menaquinone, plastoquinone) fulfill important roles in the electron transport chain [3, 6]. Enterobactin (from 2,3-dihydroxybenzoate) is used as an iron complexing siderophore in E. coli, folate is involved in C1 metabolism, and vitamin E in photosynthetic organisms is involved in protection against oxidative stress [7,8,9]. Various secondary metabolites such as phenylpropanoids (mainly in plants) and some antibiotics (chloramphenicol, pristinamycin, and others) in streptomycetes are as well derived from the aromatic pathway [10,11,12].

Aromatic amino acids are currently produced in large scale by fermentations or biotransformations with genetically modified microorganisms like E. coli or Corynebacterium glutamicum [3,4,5, 7, 13,14,15,16,17,18]. Aromatic compounds find many applications in biotechnology, nutrition, and medicine. l-Phe is mainly used for the synthesis of the low-calorie sweetener aspartame (l-aspartyl- l-phenylalanine methyl ester) while l-Trp is used for pharmaceuticals and as animal feed [3, 16]. l-Tyr is used as dietary supplement in the treatment of phenylketonuria. It is also the precursor in the synthesis of Levodopa (l-3,4-dihydroxyphenylalanine) which is used to treat Parkinson’s disease [19]. Ubiquinone-10 (Q10), folic acid, and vitamin K (plastoquinone) are used as nutritional additives [20, 21].

The rare, non-proteinogenic aromatic amino acid, para-amino-l-phenylalanine (l-PAPA) is used for technical and pharmaceutical applications [22,23,24,25] and has been described as a building block of the anticancer drug, Melphalan® [26, 27]. In nature, l-PAPA occurs in plant seeds (Vigna vexillata; [28]), and is an intermediate of the chloramphenicol and pristinamycin biosynthesis pathways in Streptomyces venezuelae and S. pristinaespiralis [29,30,31]. l-PAPA is also a precursor of the antibiotic obafluorin ß-lactone of Pseudomonas fluorescens [32] and of GameXPeptides in the entomopathogenic bacterium, Photorhabdus luminescens [33]. l-PAPA was moreover successfully used as precursor in the biosynthetic diversification of jadomycin production with S. venezuelae cells [34]. L-PAPA has also been used for the synthesis of the biopolyamide precursor, 4-aminohydrocinnamic acid [35, 36].

l-PAPA synthesis starts from chorismate which is first converted by a glutamine- or ammonia-dependent 4-amino-4-deoxychorismate synthase to give 4-amino-4-deoxychorismate (ADC) [37]. This step is common to both, the para-aminobenzoate (PABA) and the chloramphenicol/pristinamycin biosynthesis pathways. In PABA biosynthesis, ADC synthase is encoded by genes pabA and pabB—which are not organized in an operon—in E. coli [38]. In Corynebacteria, these genes are fused as pabAB such as in C. glutamicum [39,40,41], in C. efficiens [42], or C. callunae [40]. In pristinamycin biosynthesis, the pathway-specific gene is named papA in S. pristinaespiralis [30], and in chloramphenicol biosynthesis of S. venezuelae it is either termed papA [43] or cmlB [44], respectively. Despite the different nomenclature, these genes display high sequence identities [30, 43,44,45]. ADC is converted by 4-amino-4-deoxy-chorismate mutase (gene papB in S. pristinaespiralis; alternatively cmlD in S. venezuelae) and 4-amino-4-deoxyprephenate dehydrogenase (papC and cmlC, respectively) to yield p-aminophenylpyruvate (PAPP; [30, 43, 44]). By transamination, l-PAPA is formed and serves then as intermediate for the production of several antibiotics. Recently, papA,B,C homolog genes from Pseudomonas fluorescens have been successfully cloned and expressed in E. coli [24].

Escherichia coli is not a natural producer of l-PAPA but the E. coli aminotransferases TyrB and AspC are known to convert PAPP into l-PAPA (Fig. 1) [46, 47]. Earlier, l-PAPA formation in recombinant E. coli (carrying papABC genes from S. venezuelae on a low-copy plasmid) had been shown by the group of Schultz [43]. They used this in vivo produced compound as “21st proteinogenic amino acid” together with a modified tyrosyl-tRNA synthetase and a mutant tyrosine amber suppressor tRNA in specifically engineered E. coli strains. While formation of up to 0.7 mM of intracellular l-PAPA were measured in these recombinant cells, no data on extracellular production were provided [43], however.

Fig. 1
figure 1

Overview of the de novo l-PAPA biosynthesis pathway from glycerol in E. coli. The scheme of reactions is modified from Gottlieb et al. [15]. Broken arrows indicate incomplete presentation of the metabolic pathway. AroF DAHP synthase, AroB dehydroquinate synthase, AroL shikimate kinase, AspC aspartate transaminase, GlpD glycerol-3-phosphate dehydrogenase, GlpF glycerol facilitator, GlpK glycerol kinase, GlpX fructose-1,6-bisphosphate phosphatase, PabAB 4-amino-4-deoxychorismate synthase, PapB 4-amino-4-deoxychorismate mutase, PapC 4-amino-4-deoxyprephenate dehydrogenase, TktA transketolase A, TyrB aromatic aminotransferase are presented

A preferred carbon source for the production of aromatic amino acids with E. coli is glucose. Glycerol which can be used as an alternative carbon and energy source by E. coli, is especially attractive for the production of aromatic amino acids [15, 18, 48]. Glycerol can be taken up in E. coli either via the glycerol facilitator, GlpF or by unassisted diffusion [49]. After phosphorylation by the ATP-dependent glycerol kinase (GlpK; [50]), and an oxidation step by glycerol 3-phosphate dehydrogenase (GlpD; [51]) dihydroxyacetone phosphate (DHAP) can enter the lower glycolysis. Glycerol has no competing application in the food industry and it appears as a byproduct during biodiesel production [52]. Here we describe the construction of E. coli strains for the production of valuable l-PAPA using glycerol as sole carbon and energy source.

Results

Construction of a de novo l-PAPA biosynthesis pathway in E. coli

As l-PAPA is a non-proteinogenic amino acid for wildtype E. coli cells, we wanted to know first whether it shows any inhibitory or toxic effects on growth. Therefore, cells of E. coli K-12 wildtype strain LJ110 (W3110) were grown in minimal media with glycerol as sole carbon source and synthetic l-PAPA was added at varying concentrations. As can be seen in Fig. 2, the growth rate in the presence of 4 or 8 mM l-PAPA was slightly reduced (µ = 0.43 h−1 and µ = 0.39 h−1, respectively) when compared to growth in the absence of l-PAPA (µ = 0.47 h−1); however, very similar final optical densities (OD600 values) were reached after 24 h of incubation. At a concentration of 33 mM l-PAPA (5.9 g l−1), however, the formation of biomass was impaired and the growth rate was conceivably reduced to µ = 0.24 h−1. Thus we reasoned that l-PAPA production in recombinant E. coli strains could be proceeded without serious toxicity problems caused by the novel product.

Fig. 2
figure 2

Growth of E. coli LJ110 in minimal media with glycerol and different l-PAPA concentrations. The cells were cultivated at 37 °C in the absence of l-PAPA (circles), or in the presence of 4 mM (squares), 8 mM (triangles), or 33 mM of l-PAPA (diamonds), respectively. Cultivations were performed twice and the mean values are given

We decided to establish a synthetic pathway leading to l-PAPA by construction of a strain which carries the genes from different donor organisms (“metabolic grafting”). From previous work in our group [2, 41], the gene encoding 4-amino-4-deoxychorismate synthase from the PABA biosynthesis route of C. glutamicum (pabABCgl) was available in recombinant form and under the control of an IPTG-inducible Ptac promoter on plasmid pC53 (see Table 1). We preferred the pabABCgl gene over the two separate genes pabA and pabB which reside in the E. coli chromosome, as the plasmid-borne combination of the latter genes had shown less active enzyme [41]. To complete the intended l-PAPA pathway, genes papB and papC were added. PapB and papC were custom-synthesized based upon the gene sequences of S. venezuelae (papB, 4-amino-4-deoxy-chorismate mutase; papC, 4-amino-4-deoxyprephenate dehydrogenase) and codon optimized for expression in E. coli (GeneOptimizer, Thermo Fisher/GeneArt, Regensburg, Germany). The two genes were then cloned together as a cassette with pabABCgl to yield plasmid pC53BC (see Table 1, Additional file 1: Figure S4).

Table 1 List of E. coli strains, plasmids and oligonucleotides

l-PAPA production by recombinant E. coli strains

Expression of the gene cassette from pC53BC was first studied in the background of the wildtype strain E. coli LJ110. Cells were cultivated in shake flasks in minimal medium with glycerol (5 g l−1) and induction was by addition of IPTG (0.5 mM final concentration). This already led to l-PAPA accumulation of about 43 ± 2 mg l−1 in the supernatant whereas no l-PAPA was detected in the control strain (see Table 2). This proved that l-PAPA can be formed in conceivable amounts by E. coli cells and is in line with former observations with E. coli strains that were constructed in different manners [24, 43].

Table 2 Formation of l-PAPA in E. coli wild-type strain with different plasmid combinations

It is well-known that overexpression of the aroF, aroB, and aroL genes enhances formation of chorismate-derived compounds [3, 15, 59]. We therefore analyzed the effect of plasmid-borne overexpression of the genes aroF (DAHP synthase), aroB (dehydroquinate synthase) and aroL (shikimate kinase) present as a cassette on a second vector (pJNT522; KanR, see Table 1 and Additional file 1: Figure S1) which is compatible with pC53BC. The additional IPTG-induced expression of aroFBL doubled the product titer after 48 h to 86.6 ± 3.7 mg l−1 (Table 2), most likely due to an improved provision of shikimate precursors for l-PAPA formation.

Next, to avoid an undesired loss of chorismate to l-Phe and l-Tyr formation (Fig. 1), a pheA-tyrA-aroF deletion mutant of LJ110 was used as host strain. This strain (FUS 4) is a double auxotroph for l-Phe and l-Tyr; it had been successfully used previously for the production of l-Phe; in addition, strain FUS4 carries a chromosomally inserted copy of the aroFBL cassette [15]. This second generation strain showed improved l-PAPA formation in shake flasks (minimal media with glycerol and growth supplementation by l-Phe and l-Tyr; see Table 3). The l-PAPA titer was raised to about 200 mg l−1 with a yield of ca. 0.04 g l-PAPA/g of glycerol compared to E. coli LJ110 (Table 2).

Table 3 Comparison of l-PAPA yields in next generation strains

Escherichia coli FUS4.7, a FUS4 derivative which has additional chromosomal copies of tktA (transketolase A) and glpX (fructose 1,6-bisphosphate phosphatase), respectively, had turned up as an improved l-Phe producer before [15]. It was therefore used as recipient for pC53BC and pJNT-aroFBL to study its l-PAPA productivity. Indeed, the titer further increased to 449 ± 29 mg l−1 and the yield was more than double in comparison to FUS 4 with the same plasmid combination (see Table 3).

Finally, the gene for the transcriptional regulator, tyrR [60], was deleted. Such a deletion had been shown to confer a positive effect for aromatic compound production in engineered E. coli strains [61,62,63,64,65,66]. Deletion of tyrR in FUS 4.7 led to strain FUS4.7R. When transformed with plasmids pC53BC and pJNT-aroFBL this combination resulted in a slightly increased titer of l-PAPA (534 ± 24 mg l−1) and a yield of ca. 11% l-PAPA/glycerol (g g−1) (Table 3). The biomass yield of the three strains in shake flasks was about the same.

Fed-batch shake flask cultivation with E. coli FUS4.7R/pC53BC/pJNT-aroFBL

These encouraging results from shake flask cultivations prompted us to study growth behavior and l-PAPA production in fed-batch condition. We chose the best producer, so far, E. coli FUS4.7R/pC53BC/pJNT-aroFBL. After pre-incubation and appropriate IPTG induction in shake flask with a starting concentration of 5 g l−1 of glycerol at 37 °C, temperature was lowered to 30 °C and glycerol (~ 5 g l−1) was added in intervals of 12 h. To roughly adjust for the pH, sodium bicarbonate was added (30 mM) every 24 h, as well as 15 mM ammonium sulfate as a nitrogen source to allow formation of l-PAPA. As a result, a total of 33.3 g l−1 glycerol was consumed in 134 h of cultivation (Fig. 3) and a titer of 5.47 ± 0.41 g l−1 of l-PAPA was detected in the supernatant; this corresponds to a yield of 16% l-PAPA/glycerol (g g−1).

Fig. 3
figure 3

Glycerol fed-batch cultivation of E. coli FUS4.7 R/pC53BC/pJNT-aroFBL in shake flasks. The initial glycerol concentration was 5 g l−1. After 24 h of cultivation the cultures were induced with 0.5 mM IPTG (final concentration) and the shake flasks were transferred from 37 to 30 °C before adding pulses of ~ 5 g l−1 of glycerol every 12 h. To adjust the pH, sodium bicarbonate (30 mM) was added every 24 h as well as 15 mM of ammonium sulfate. The concentrations of glycerol (empty circles) and, l-PAPA (empty squares) were determined by HPLC. OD600 values are presented as filled squares. The cultivations were performed in triplicate and the mean values and standard deviations are given

Fed-batch fermentation in a bioreactor

In order to scale up L-PAPA production under controlled conditions, we cultivated E. coli FUS4.7R/pC53BC/pJNT-aroFBL cells in an aerated, pH-controlled 30 l stirred-tank reactor with minimal medium and glycerol as sole carbon source. The starting volume for the batch was 8 l and the final volume (after fed-batch) of 12.25 l. Details of the fermentation procedure are given in the Materials and Methods section. The feeding started after the initially added glycerol (8.1 g l−1) was consumed. Glycerol was then fed to maintain a concentration between 0.6 and 1.0 g l−1. Additionally, l-Phe and l-Tyr were added in two pulses to allow biomass formation. After 77 h, a final cell dry weight of 21.6 g l−1 was reached and 131.7 g l−1 of glycerol had been consumed (see Fig. 4). A total concentration of 16.78 g l−1 l-PAPA was produced by then. This equals a yield of 0.13 l-PAPA/glycerol (g g−1) and a space–time-yield of l-PAPA formation over the whole process of 0.22 g l−1 h−1. The total amount of l-PAPA produced during the 77 h of fermentation was 205 g. The produced l-PAPA was analyzed by MS and the obtained mass was in excellent agreement with the commercially available l-PAPA (Additional file 2).

Fig. 4
figure 4

Fed-batch production of l-PAPA in a 30 l bioreactor. Cell dry weight (CDW, empty circle), glycerol consumption (g l−1, empty triangle) and l-PAPA concentration (g l−1, empty square) were monitored for E. coli FUS4.7R/pC53BC/pJNT-aroFBL. The strain was grown in minimal media with an initial concentration of 8.1 g l−1 glycerol. The glycerol feed started after 11.4 h (red solid line, first pulse of l-Phe and l-Tyr and induction with 0.5 mM IPTG) and the feed was adjusted to maintain a concentration of 0.6–1.0 g l−1 glycerol. A second pulse of l-Phe and l-Tyr (dotted line) and a third pulse of l-Phe and l-Tyr (dash line) were fed to allow biomass formation. The result of one fermentation is presented with mean values and deviations of three technical replicates

To our knowledge, this is the highest l-PAPA titer which has been reached with E. coli cells.

Discussion

Expansion of the chorismate pathway of E. coli to produce a variety of non-standard aromatic compounds has already been very successful in many cases. Pioneering work has been done by the group of Frost and others already in the 1990s [61,62,63,64, 66]. Thereafter, processes have been published for p-hydroxybenzoic acid [61, 65], p-aminobenzoic acid [40, 42], protocatechuic acid and catechol [65], vanillin [67, 68], anthranilic acid [69], δ-tocotrienol (a vitamin E compound, [70]), phenol [71,72,73,74], salicylic acid and muconic acid [5, 75], styrene, cinnamic acid and hydroxylated derivatives thereof [7, 76,77,78,79,80], tyrosol [81], (hydroxyl)phenyllactic acid, tyramine and other l-Phe- or l-Tyr-derived compounds [76], rosmarinic acid and flavonoids [82, 83], antitumor drugs like violacein and deoxyviolacein [84], several plant alkaloids [85], or even opiates like thebaine, reticuline or hydrocodone [86]. Further examples may be found in recent review articles [79, 87, 88].

These approaches most often have in common that heterologous genes from other donor organisms and from other biosynthetic pathways are used for transplantation into the genome of recipient strains such as E. coli. These recipients often have been engineered for a fortified provision of chorismate from the general aromatic amino acid pathway [61, 66, 69, 89]. We propose to classify this subdivision of “metabolic engineering” [90] as “metabolic grafting” [3] as it may be metaphorically likened to the well-known and centuries old procedure in plant husbandry which is termed “grafting”. Therein, for example, a fledgling twig from a precious fruit tree is merged (pruned) with a robust stem of another tree (which may be of another plant species), resulting in strong fruit bearing by the grafted-upon plant (see for example the case of using North American vine as rootstock for many rare and precious European wines to overcome the Phylloxera vine pest in the late nineteenth century [91]). In analogy, the transfer of such a grafting procedure to the expansion of the chorismate pathway in microorganisms thus implies cloning and expression of heterologous genes to allow production of non-standard products. Whereas the proteinogenic aromatic amino acids l-Phe, l-Tyr, and l-Trp can be produced by recombinant E. coli strains even without the introduction of foreign genes ([3, 89, 92], non-standard and non-proteinogenic aromatic amino acids require such a grafting procedure. Examples are the production of l-homophenylalanine by introduction of genes from the cyanobacterium Nostoc punctiforme for the chain elongation of l-Phe [93], improved production of l-3,4-dihydroxyphenylalanine (l-DOPA) from E. coli grown on glucose [94, 95], or the production of d-phenylglycine via phenylpyruvic acid and mandelic acid [96].

Following the initial work of the Schultz group on l-PAPA formation in recombinant E. coli strains using the genes papABC from S. venezuelae [43], Takaya and coworkers recently presented a genetically modified E. coli to produce l-PAPA from glucose by introduction of papABC genes from P. fluorescens [24]. A high titer of 4.4 l-PAPA g l−1 with a production yield of 17% (g g−1) from glucose was achieved in minimal media supplemented with tryptone (5 g l−1) and yeast extract (2.5 g l−1) in a bioreactor [24].

The aim of the present study was to produce the high value compound l-PAPA with genetically engineered E. coli strains from glycerol, a renewable carbon source (for a recent review see [48]). Glycerol has become a favorable carbon source as it is a byproduct of biodiesel production [15, 48, 52, 97]. The advantage to use glycerol instead of glucose is that there is no need of stoichiometric amounts of phosphoenolpyruvate (PEP) for its uptake into E. coli cells, whereas the PTS- dependent glucose uptake needs PEP [98]. PEP and E4P are important precursors of the initial reaction of shikimate pathway to form DAHP by the DAHP synthase [99]. Furthermore, glycerol (a C3 unit), compared to glucose (a C6 unit), has a higher reduction potential per C3 unit which in turn could yield higher biomass and product formation [15, 48, 100,101,102]. The precursor supply is a pivotal parameter to improve production of compound with origin in the shikimate pathway [17]. Our group had shown in previous studies [15, 18] that the improved supply of E4P by integration of additional gene copies of the fructose 1,6-bisphosphate phosphatase (glpX) and transketolase (tktA) increased the yield for l-Phe production [15, 18]. Therefore we applied this knowledge here and found that l-PAPA synthesis also profited from these genetic alterations. The extra gene copies in FUS 4.7 led to a doubled product titer of about 0.45 ± 0.03 g l−1 in batch culture shake flasks after 48 h compared to the parent strain FUS4 (0.20 ± 0.01 g l−1) in minimal medium with 5 g l−1 glycerol (Table 2).

Moreover, we analyzed the effect of a tyrR deletion in the recombinant E. coli FUS4.7. The transcriptional regulon of TyrR is well characterized in E. coli [60]. TyrR downregulates as transcriptional repressor the genes aroF, aroG, aroL, tyrP, aroP, tyrA and tyrB in the presence of l-Tyr [60, 103]. We observed a beneficial effect on l-PAPA production by a tyrR deletion in E. coli FUS4.7R compared to E. coli FUS4.7. The yield was increased by ~ 18% from 0.09 l-PAPA/glycerol (g g−1) to 0.11 l-PAPA/glycerol (g g−1) in batch cultivation. This result is in good agreement with previous studies [76, 92, 94]. It is likely that the positive effect of a tyrR deletion relies on the increased transcriptional expression and resulting increased specific activities for members of the TyrR regulon.

By changing the cultivation condition of E. coli FUS4.7R to a fed-batch cultivation in shake flasks, we improved the total yield to 0.16 l-PAPA/glycerol (g g−1) and reached a titer of about 5.47 ± 0.41 g l−1 l-PAPA after 134 h. This titer is already higher than the 4.4 g l−1 obtained found by the Takaya group in minimal media with glucose as main C source (plus added tryptone and yeast extract) [24]. However, as the cultivation in fed-batch shake flasks is not optimal in term of oxygen supply and pH stability it may be advantageous to perform the cultivation in a pH- and pO2-controlled bioreactor. Indeed, we found that during a run of 77 h, 131.7 g l−1 of glycerol were consumed and a total of 205 g of l-PAPA were produced (16.78 g l−1). Only small amounts of acetate (< 2 g l−1) and no lactate as by-products were detected during fermentation. This corresponds to a yield of 0.13 g l-PAPA per g of glycerol, slightly lower than in the fed-batch shake flask cultures. The reached titer in the bioreactor showed that l-PAPA production at higher concentration is still feasible although the growth of E. coli LJ110 was impaired at 5.9 g l−1 (Fig. 2). In comparison with other products of “extended shikimate pathway” [7], the presented data on l-PAPA are in the same range as p-aminobenzoate from glucose in recombinant E. coli (4.8 g l−1 in 48 h; yield of 0.16 g g−1 [79]) but below anthranilate (o-aminobenzoate: 14 g l−1 in 34 h on glucose with a yield of 0.2 g g−1 [69]) where both products, however, need only one ammonium equivalent for their production.

To further improve the E4P precursor supply it can be considered to further increase the number of gene copies in the E. coli genome as only one additional copy of tktA and glpX was inserted. The enhanced flux through the shikimate pathway by overexpression of plasmid-borne aroFBL genes showed a beneficial effect in our study. Other studies have shown that the overexpression of the other shikimate pathway genes (aroA, C, D and E) as well improved the formation of l-Tyr which is derived from chorismate [14]. It is likely that a positive effect can be also observed for the l-PAPA production. This improved route to l-PAPA based upon a simple medium with a sustainable carbon source which is currently in ample supply from biodiesel production and which does not compete with human nutrition, opens up the possibility to produce an interesting building block in larger scale.

Conclusion

This study demonstrated that E. coli is a good chassis strain for l-PAPA production using glycerol as an alternative sole carbon source. We constructed by metabolic grafting a de novo pathway for l-PAPA in E. coli. By improving the E4P precursor supply and the increased flux through the shikimate pathway both l-PAPA titer and yield were augmented.

Methods

Chemicals, antibiotics, buffer components, culture media and analytical standards, used in this study were purchased from AppliChem GmbH (Darmstadt, Germany), Carl Roth GmbH (Karlsruhe, Germany), or Sigma-Aldrich/Fluka (Taufkirchen, Germany) and were of the highest available purity. l-PAPA was purchased from Sigma-Aldrich Chemie GmbH (Taufkirchen, Germany).

Bacterial strains, plasmids and cultivation conditions

All bacterial strains, plasmids, and oligonucleotides used in this study are listed in Table 1. For the cultivation of E. coli, lysogeny broth (LB) was used as complex media [104]. The minimal media (MM) contained 3 g l−1 KH2PO4, 12 g l−1 K2HPO4, 5 g l−1 (NH4)2SO4, 0,3 g l−1 MgSO4·7H2O, 0,1 g l−1 NaCl, 0,1125 g l−1 FeSO4·7H2O/Na citrate 15 ml (from the solution of 7.5 g l−1 FeSO4 and 100 g l−1 sodium citrate), 0,015 g l−1 CaCl2·2H2O, 7.5 µg l−1 thiamine HCl, 0,04 mg ml−1 l-Phe, 0,04 mg ml−1 l-Tyr and 5 g l−1 glycerol as sole carbon source [15]. Ampicillin sodium salt (100 mg l−1) and/or kanamycin sulfate (50 mg l−1) was added when appropriate.

Escherichia coli strains were grown in 250 ml shake flasks filled with 20 ml of medium at 37 °C and with 150-rpm agitation. Overnight cultures were used as inocula (1% v/v). Induction was with 0.5 mM isopropyl-β-d-thiogalactopyranoside (IPTG, final concentration) at an OD at 600 nm (OD600) of about 0.6. Samples were taken and centrifuged at 22,000g for 10 min. The supernatants were removed and stored at − 20 °C until further use. The glycerol fed-batch cultivation was performed in 250 ml shake flasks with 20 ml media. The initial glycerol concentration was 5 g l−1. Cultures were inoculated with an overnight grown culture. After 24 h of cultivation, 0.5 mM IPTG (final concentration) was added to the cells and the shake flasks were transferred from 37 to 30 °C before starting to add ~ 5 g l−1 glycerol to the culture every 12 h until 84 h. Furthermore, every 24 h, 30 mM sodium bicarbonate was added to adjust the pH and 15 mM ammonium sulfate was added as ammonium source. Samples were taken and centrifuged at 22,000g for 10 min. The supernatants were removed and stored at − 20 °C until use.

Fed-batch fermentation in a 30 l bioreactor

A 30 l stirred-tank reactor (Bioengineering, Wald, Switzerland) with a batch volume of 8 l and a final volume of 12.25 l was used and a minimal medium with glycerol as sole carbon source. The concentration of supplementation compounds was increased 1.5-fold, e.g. supplementation was with 60 mg l−1 l-Phe, 60 mg l−1 of l-Tyr. The concentration of antibiotics was kept at 100 mg l−1 ampicillin sodium salt and 50 mg l−1 kanamycin sulfate. The working conditions were pH 7.0 (controlled by the addition of 15% ammonia solution), 30% oxygen saturation with a reactor pressure of 500 hPa above atmospheric pressure (starting with pO2 ~ 100% at 350 rpm with an aeration rate of 4 l min−1; when reaching 30% oxygen saturation, the stirring rate was adjusted). A preculture (1.2 l grown in shake flasks with the same medium) which was in the exponential growth in shake flasks, was used to inoculate the fermentation (starting CDW ~ 0.18 g l−1) t 37 °C. 11.75 h after inoculation IPTG (0.5 mM), l-Tyr (0.36 mM) and l-Phe (0.33 mM) were added and the temperature was shifted from 37 to 30 °C. The initial glycerol concentration was about 8.1 g l−1. The glycerol feed was started after the initial glycerol was depleted (indicated by a rise in pO2). Glycerol was fed to maintain a concentration between 0.6 and 1.0 g l−1 during cultivation. If needed, antifoam agent (Struktol J647) from Schill Seilacher (Hamburg) was added to the culture. Additionally, l-Phe and l-Tyr were added in two pulses (34 h after inoculation 0.5 mM l-Phe and l-Tyr; 56 h after inoculation 0.3 mM l-Phe and l-Tyr) to allow biomass formation.

DNA manipulations

For plasmid constructions, E. coli strain DH5α was used. Standard molecular biology methods were applied [104]. DNA containing the genes papBC from S. venezuelae was custom synthesized (Thermo Fisher/GeneArt, Regensburg, Germany) and codon optimized for E. coli to give pMK-papBC (see Table 1); this plasmid DNA was then restricted with BglII and BamHI. A resulting fragment containing papBC was isolated and ligated with a BamHI restricted pC53 vector (based on pJF119 EH vector [44] which already carries the pabAB genes from C. glutamicum [41]) to obtain pC53BC (see Additional file 1: Figure S4). The recombinant genes are under the control of the lacIq/Ptac promoter. The sequence and direction of papBC was verified by sequencing (GATC Biotech AG, Konstanz, Germany). Plasmid pJNT-aroFBL is based on vector pJNT522 which in turn is derived from pJeM2 [58] and pJF119EH. Construction of pJNT522 vector is described in the Additional file 1. In brief, pJNT522 contains a fragment of pJeM2 plasmid carrying the mob region, origin of replication, and kanamycin resistance gene as well as the Ptac promoter, MCS and rrnB T1T2 terminator of transcription from the pJF119EH vector. pJNT522 is compatible with pJF119EH- derived plasmids. The subcloning of aroFB and aroL genes of E. coli and ligation on vector pJNT522 to yield pJNT-aroFBL is described in the Additional file 1.

Deletion of tyrR gene from the E. coli chromosome

The deletion of the gene tyrR, encoding the regulator of the tyr regulon, was carried out according to a λred-recombineering method [55]. A linear DNA fragment containing the FRT-flanked chloramphenicol resistance (cat) cassette was amplified from plasmid pCAS30-FRT-cat-FRT [57] using the primer pair DeltyrRFw/-Rw (see Table 1 for sequences). The thus obtained amplified linear DNA was introduced by electroporation into electrocompetent E. coli FUS4.7 cells that carried the Red recombinase expression vector pKD46 [55]. After confirmation of the FRT-cat-FRT integration by colony PCR with the primer pair Ko-tyrRFw/-Rw (see Table 1), the cat marker was removed by transient expression of a FLP recombinase from plasmid pCP20 [56] to eventually generate the tyrR deletion strain, FUS4.7R. Verification of the disruption was performed using colony PCR with primers up- and downstream of disrupted regions (Ko-tyrR-Fw/-Rw). Finally, cells were grown at 42 °C to remove the temperature-sensitive replicons, pCP20 or pKD46, respectively [55].

Analytical methods

Growth of E. coli cells was followed by measuring OD600 in a UV–vis spectrophotometer (Cary 50 Bio, Agilent Technologies). The formation of l-PAPA was routinely analyzed and quantified by high-performance liquid chromatography (HPLC, 1260 Infinity series, Agilent Technologies) with a diode array detector (1260 Infinity series, Agilent Technologies) at a wavelength of 210 nm. A Prontosil C18 column (250 × 4 mm, CS-Chromatographie Services GmbH, Langerwehe, Germany) was used for separation at 40 °C. The mobile phase containing 40 mM Na2SO4 (adjusted at pH ~ 2.7 with methane sulfonic acid) was used with a flow rate of 1 ml min−1. Glycerol concentrations were determined by HPLC with a refractive index detector (1260 Infinity series, Agilent Technologies) and an Organic Acid column (300 × 8 mm, CS-Chromatographie Services GmbH, Langerwehe Germany). An isocratic flow of 0.6 ml min −1 of 5 mM H2SO4 was applied at 40 °C. The produced L-PAPA was analyzed by mass spectrometry. It was determined by using an Agilent 6130 mass spectrometer system with electron spray ionization (Agilent Technologies; Germany).

Abbreviations

ADC:

4-amino-4-deoxychorismate

ADP:

adenosine diphosphate

α-KG:

alpha-ketoglutarate

Amp:

ampicillinAspC aspartate aminotransferase

AroB:

dehydroquinate synthase

AroF:

DAHP synthase

AroL:

shikimate kinase

ATP:

adenosine triphosphate

DAHP:

3-deoxy-d-arabino-heptulosonate 7-phosphate

DHAP:

dihydroxyacetone phosphate

DHQ:

dehydroquinate

E4P:

erythrose-4-phosphate

F6P:

fructose-6-phosphate

FBP:

fructose-1,6-bisphosphate

G3P:

glycerol-3-phosphate

GAP:

glyceraldehyde-3-phosphate

GlpD:

glycerol-3-phosphate dehydrogenase

GlpF:

glycerol facilitator

GlpK:

glycerol kinase

Glu:

l-glutamate

Gln:

l-glutamine

GlpX:

fructose-1,6-bisphosphatase

IPTG:

isopropyl-β-d-thiogalactopyranoside

Kan:

kanamycin

l-PAPA:

para-amino-l-phenylalanine

l-Phe:

l-phenylalanine

l-Trp:

l-typtophan

l-Tyr:

l-tyrosine

MQ:

menaquinone

OD:

optical density

PABA:

para-aminobenzoate

PabAB:

4-amino-4-deoxychorismate synthase

PapB:

4-amino-4-deoxychorismate mutase

PapC:

4-amino-4-deoxyprephenate dehydrogenase

PAPP:

para-aminophenylpyruvate

PEP:

phosphoenolpyruvate

PTS:

PEP-dependent sugar:phosphotransferase system

TCA:

tricarboxylic acid cycle

TktA:

transketolase A

TyrB:

aromatic aminotransferase

UQ:

ubiquinone

References

  1. Gibson F, Jackman LM. Structure of chorismic acid, a new intermediate in aromatic biosynthesis. Nature. 1963;198:388–9.

    Article  CAS  PubMed  Google Scholar 

  2. Bongaerts J, Esser S, Lorbach V, Al-Momani L, Muller MA, Franke D, Grondal C, Kurutsch A, Bujnicki R, Takors R, et al. Diversity-oriented production of metabolites derived from chorismate and their use in organic synthesis. Angew Chem Int Ed Engl. 2011;50:7781–6.

    Article  CAS  PubMed  Google Scholar 

  3. Sprenger GA. From scratch to value: engineering Escherichia coli wild type cells to the production of l-phenylalanine and other fine chemicals derived from chorismate. Appl Microbiol Biotechnol. 2007;75:739–49.

    Article  CAS  PubMed  Google Scholar 

  4. Berry A. Improving production of aromatic compounds in Escherichia coli by metabolic engineering. Trends Biotechnol. 1996;14:250–6.

    Article  CAS  PubMed  Google Scholar 

  5. Noda S, Shirai T, Oyama S, Kondo A. Metabolic design of a platform Escherichia coli strain producing various chorismate derivatives. Metab Eng. 2016;33:119–29.

    Article  CAS  PubMed  Google Scholar 

  6. van Beilen JWA, Hellingwerf KJ. All three endogenous quinone species of Escherichia coli are involved in controlling the activity of the aerobic/anaerobic response regulator ArcA. Front Microbiol. 2016;7:1339.

    PubMed  PubMed Central  Google Scholar 

  7. Lee JH, Wendisch VF. Biotechnological production of aromatic compounds of the extended shikimate pathway from renewable biomass. J Biotechnol. 2017;257:211–21.

    Article  CAS  PubMed  Google Scholar 

  8. Young IG, Langman L, Luke RK, Gibson F. Biosynthesis of the iron-transport compound enterochelin: mutants of Escherichia coli unable to synthesize 2,3-dihydroxybenzoate. J Bacteriol. 1971;106:51–7.

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Mene-Saffrane L, DellaPenna D. Biosynthesis, regulation and functions of tocochromanols in plants. Plant Physiol Biochem. 2010;48:301–9.

    Article  CAS  PubMed  Google Scholar 

  10. Mast Y, Weber T, Golz M, Ort-Winklbauer R, Gondran A, Wohlleben W, Schinko E. Characterization of the ‘pristinamycin supercluster’ of Streptomyces pristinaespiralis. Microb Biotechnol. 2011;4:192–206.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Lin DR, Xiao MS, Zhao JJ, Li ZH, Xing BS, Li XD, Kong MZ, Li LY, Zhang Q, Liu YW, et al. An overview of plant phenolic compounds and their importance in human nutrition and management of type 2 diabetes. Molecules. 2016;21:1374.

    Article  PubMed Central  Google Scholar 

  12. Fernandez-Martinez LT, Borsetto C, Gomez-Escribano JP, Bibb MJ, Al-Bassam MM, Chandra G, Bibb MJ. New insights into chloramphenicol biosynthesis in Streptomyces venezuelae ATCC 10712. Antimicrob Agents Chemother. 2014;58:7441–50.

    Article  PubMed  PubMed Central  Google Scholar 

  13. Schuhmacher T, Loffler M, Hurler T, Takors R. Phosphate limited fed-batch processes: impact on carbon usage and energy metabolism in Escherichia coli. J Biotechnol. 2014;190:96–104.

    Article  CAS  PubMed  Google Scholar 

  14. Juminaga D, Baidoo EE, Redding-Johanson AM, Batth TS, Burd H, Mukhopadhyay A, Petzold CJ, Keasling JD. Modular engineering of l-tyrosine production in Escherichia coli. Appl Environ Microbiol. 2012;78:89–98.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Gottlieb K, Albermann C, Sprenger GA. Improvement of l-phenylalanine production from glycerol by recombinant Escherichia coli strains: the role of extra copies of glpK, glpX, and tktA genes. Microb Cell Fact. 2014;13:96.

    Article  PubMed  PubMed Central  Google Scholar 

  16. Ikeda M. Towards bacterial strains overproducing l-tryptophan and other aromatics by metabolic engineering. Appl Microbiol Biotechnol. 2006;69:615–26.

    Article  CAS  PubMed  Google Scholar 

  17. Rodriguez A, Martinez JA, Flores N, Escalante A, Gosset G, Bolivar F. Engineering Escherichia coli to overproduce aromatic amino acids and derived compounds. Microb Cell Fact. 2014;13:126.

    PubMed  PubMed Central  Google Scholar 

  18. Weiner M, Trondle J, Albermann C, Sprenger GA, Weuster-Botz D. Metabolic control analysis of l-phenylalanine production from glycerol with engineered E. coli using data from short-term steady-state perturbation experiments. Biochem Eng J. 2017;126:86–100.

    Article  CAS  Google Scholar 

  19. Min K, Park K, Park DH, Yoo YJ. Overview on the biotechnological production of l-DOPA. Appl Microbiol Biotechnol. 2015;99:575–84.

    Article  CAS  PubMed  Google Scholar 

  20. Cluis CP, Ekins A, Narcross L, Jiang H, Gold ND, Burja AM, Martin VJ. Identification of bottlenecks in Escherichia coli engineered for the production of CoQ(10). Metab Eng. 2011;13:733–44.

    Article  CAS  PubMed  Google Scholar 

  21. Xu W, Yang S, Zhao J, Su T, Zhao L, Liu J. Improving coenzyme Q8 production in Escherichia coli employing multiple strategies. J Ind Microbiol Biotechnol. 2014;41:1297–303.

    Article  CAS  PubMed  Google Scholar 

  22. Kumar A, Tateyama S, Yasaki K, Ali MA, Takaya N, Singh R, Kaneko T. (1)H NMR and FT-IR dataset based structural investigation of poly(amic acid)s and polyimides from 4,4′-diaminostilbene. Data Brief. 2016;7:123–8.

    Article  PubMed  PubMed Central  Google Scholar 

  23. Yamaguchi M, Saito R, Utsumi K, Ochiai A, Kawashima N, Tokuoka Y, Katoh A. The nicotinic acid-p-aminophenylalanine-hydroxybenzoic acid triads induce apoptosis in human leukemia U937 cells. Heterocycles. 2007;71:1503–8.

    Article  CAS  Google Scholar 

  24. Masuo S, Zhou S, Kaneko T, Takaya N. Bacterial fermentation platform for producing artificial aromatic amines. Sci Rep. 2016;6:25764.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Suvannasara P, Tateyama S, Miyasato A, Matsumura K, Shimoda T, Ito T, Yamagata Y, Fujita T, Takaye N, Kaneko T. Biobased polyimides from 4-aminocinnamic acid photodimer. Macromolecules. 2014;47:1586–93.

    Article  CAS  Google Scholar 

  26. Bergel F, Stock JA. Cyto active amino acid and peptide derivatives. I. Substituted phenylalanines. J Chem Soc. 1954. https://doi.org/10.1039/JR9540002409.

    Article  Google Scholar 

  27. Jobdevairakkam C, Velladurai H. Process of making optically pure melphalan. WO2009117164A1. 2013.

  28. Dardenne GA, Larsen PO, Wieczorkowska E. Biosynthesis of p-aminophenylalanine: part of a general scheme for the biosynthesis of chorismic acid derivatives. Biochim Biophys Acta. 1975;381:416–23.

    Article  CAS  PubMed  Google Scholar 

  29. Chang Z, Sun Y, He J, Vining LC. p-Aminobenzoic acid and chloramphenicol biosynthesis in Streptomyces venezuelae: gene sets for a key enzyme, 4-amino-4-deoxychorismate synthase. Microbiology. 2001;147:2113–26.

    Article  CAS  PubMed  Google Scholar 

  30. Blanc V, Gil P, Bamas-Jacques N, Lorenzon S, Zagorec M, Schleuniger J, Bisch D, Blanche F, Debussche L, Crouzet J, Thibaut D. Identification and analysis of genes from Streptomyces pristinaespiralis encoding enzymes involved in the biosynthesis of the 4-dimethylamino-l-phenylalanine precursor of pristinamycin I. Mol Microbiol. 1997;23:191–202.

    Article  CAS  PubMed  Google Scholar 

  31. Mast YJ, Wohlleben W, Schinko E. Identification and functional characterization of phenylglycine biosynthetic genes involved in pristinamycin biosynthesis in Streptomyces pristinaespiralis. J Biotechnol. 2011;155:63–7.

    Article  CAS  PubMed  Google Scholar 

  32. Schaffer JE, Reck MR, Prasad NK, Wencewicz TA. beta-Lactone formation during product release from a nonribosomal peptide synthetase. Nat Chem Biol. 2017;13:737–44.

    Article  CAS  PubMed  Google Scholar 

  33. Nollmann FI, Dauth C, Mulley G, Kegler C, Kaiser M, Waterfield NR, Bode HB. Insect-specific production of new GameXPeptides in Photorhabdus luminescens TTO1, widespread natural products in entomopathogenic bacteria. ChemBioChem. 2015;16:205–8.

    Article  CAS  PubMed  Google Scholar 

  34. Martinez-Farina CF, Robertson AW, Yin HM, Monro S, McFarland SA, Syvitski RT, Jakeman DL. Isolation and synthetic diversification of jadomycin 4-amino-l-phenylalanine. J Nat Prod. 2015;78:1208–14.

    Article  CAS  PubMed  Google Scholar 

  35. Kawasaki Y, Aniruddha N, Minakawa H, Masuo S, Kaneko T, Takaya N. Novel polycondensed biopolyamide generated from biomass-derived 4-aminohydrocinnamic acid. Appl Microbiol Biotechnol. 2018;102:631–9.

    Article  CAS  PubMed  Google Scholar 

  36. Tateyama S, Masuo S, Suvannasara P, Oka Y, Miyazato A, Yasaki K, Teerawatananond T, Muangsin N, Zhou SM, Kawasaki Y, et al. Ultrastrong, transparent polytruxillamides derived from microbial photodimers. Macromolecules. 2016;49:3336–42.

    Article  CAS  Google Scholar 

  37. Viswanathan VK, Green JM, Nichols BP. Kinetic characterization of 4-amino 4-deoxychorismate synthase from Escherichia coli. J Bacteriol. 1995;177:5918–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Nichols BP, Seibold AM, Doktor SZ. para-aminobenzoate synthesis from chorismate occurs in two steps. J Biol Chem. 1989;264:8597–601.

    CAS  PubMed  Google Scholar 

  39. Stolz M, Peters-Wendisch P, Etterich H, Gerharz T, Faurie R, Sahm H, Fersterra H, Eggeling L. Reduced folate supply as a key to enhanced l-serine production by Corynebacterium glutamicum. Appl Environ Microbiol. 2007;73:750–5.

    Article  CAS  PubMed  Google Scholar 

  40. Kubota T, Watanabe A, Suda M, Kogure T, Hiraga K, Inui M. Production of para-aminobenzoate by genetically engineered Corynebacterium glutamicum and non-biological formation of an N-glucosyl byproduct. Metab Eng. 2016;38:322–30.

    Article  CAS  PubMed  Google Scholar 

  41. Kozak S. Mikrobielle Biosynthese von (3S,4R)-4-amino-3-hydroxycyclohexa-1,5-diencarbonsäure mit rekombinanten Escherichia coli Zellen: Molekulargenetische und biochemische Untersuchungen. Ph.D. Thesis, University of Stuttgart, Germany; 2006.

  42. Koma D, Yamanaka H, Moriyoshi K, Sakai K, Masuda T, Sato Y, Toida K, Ohmoto T. Production of p-aminobenzoic acid by metabolically engineered Escherichia coli. Biosci Biotechnol Biochem. 2014;78:350–7.

    Article  CAS  PubMed  Google Scholar 

  43. Mehl RA, Anderson JC, Santoro SW, Wang L, Martin AB, King DS, Horn DM, Schultz PG. Generation of a bacterium with a 21 amino acid genetic code. J Am Chem Soc. 2003;125:935–9.

    Article  CAS  PubMed  Google Scholar 

  44. He J, Magarvey N, Piraee M, Vining LC. The gene cluster for chloramphenicol biosynthesis in Streptomyces venezuelae ISP5230 includes novel shikimate pathway homologues and a monomodular non-ribosomal peptide synthetase gene. Microbiology. 2001;147:2817–29.

    Article  CAS  PubMed  Google Scholar 

  45. Yanai K, Sumida N, Okakura K, Moriya T, Watanabe M, Murakami T. Para-position derivatives of fungal anthelmintic cyclodepsipeptides engineered with Streptomyces venezuelae antibiotic biosynthetic genes. Nat Biotechnol. 2004;22:848–55.

    Article  CAS  PubMed  Google Scholar 

  46. Hayashi H, Inoue K, Nagata T, Kuramitsu S, Kagamiyama H. Escherichia coli aromatic amino acid aminotransferase: characterization and comparison with aspartate aminotransferase. Biochemistry. 1993;32:12229–39.

    Article  CAS  PubMed  Google Scholar 

  47. Onuffer JJ, Ton BT, Klement I, Kirsch JF. The use of natural and unnatural amino acid substrates to define the substrate specificity differences of Escherichia coli aspartate and tyrosine aminotransferases. Protein Sci. 1995;4:1743–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Sprenger GA. Glycerol as carbon source for production of added-value compounds. In: Gosset G, editor. Engineering of microorganisms for the production of chemicals and biofuels from renewable resources. Cham: Springer International Publishing AG; 2017. p. 93–123. ISBN 978-3-319-51728-5.

    Chapter  Google Scholar 

  49. Richey DP, Lin EC. Importance of facilitated diffusion for effective utilization of glycerol by Escherichia coli. J Bacteriol. 1972;112:784–90.

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Zwaig N, Kistler WS, Lin EC. Glycerol kinase, the pacemaker for the dissimilation of glycerol in Escherichia coli. J Bacteriol. 1970;102:753–9.

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Weiner JH, Heppel LA. Purification of the membrane-bound and pyridine nucleotide-independent l-glycerol 3-phosphate dehydrogenase from Escherichia coli. Biochem Biophys Res Commun. 1972;47:1360–5.

    Article  CAS  PubMed  Google Scholar 

  52. Almeida JR, Favaro LC, Quirino BF. Biodiesel biorefinery: opportunities and challenges for microbial production of fuels and chemicals from glycerol waste. Biotechnol Biofuels. 2012;5:48.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Bachmann BJ. Pedigrees of some mutant strains of Escherichia coli K-12. Bacteriol Rev. 1972;36:525–57.

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Furste JP, Pansegrau W, Frank R, Blocker H, Scholz P, Bagdasarian M, Lanka E. Molecular cloning of the plasmid RP4 primase region in a multi-host-range tacP expression vector. Gene. 1986;48:119–31.

    Article  CAS  PubMed  Google Scholar 

  55. Datsenko KA, Wanner BL. One-step inactivation of chromosomal genes in Escherichia coli K-12 using PCR products. Proc Natl Acad Sci USA. 2000;97:6640–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Cherepanov PP, Wackernagel W. Gene disruption in Escherichia coli: TcR and KmR cassettes with the option of Flp-catalyzed excision of the antibiotic-resistance determinant. Gene. 1995;158:9–14.

    Article  CAS  PubMed  Google Scholar 

  57. Vallon T, Ghanegaonkar S, Vielhauer O, Muller A, Albermann C, Sprenger G, Reuss M, Lemuth K. Quantitative analysis of isoprenoid diphosphate intermediates in recombinant and wild-type Escherichia coli strains. Appl Microbiol Biotechnol. 2008;81:175–82.

    Article  CAS  PubMed  Google Scholar 

  58. Jeske M, Altenbuchner J. The Escherichia coli rhamnose promoter rhaP(BAD) is in Pseudomonas putida KT2440 independent of Crp-cAMP activation. Appl Microbiol Biotechnol. 2010;85:1923–33.

    Article  CAS  PubMed  Google Scholar 

  59. Gosset G. Production of aromatic compounds in bacteria. Curr Opin Biotechnol. 2009;20:651–8.

    Article  CAS  PubMed  Google Scholar 

  60. Pittard J, Camakaris H, Yang J. The TyrR regulon. Mol Microbiol. 2005;55:16–26.

    Article  CAS  PubMed  Google Scholar 

  61. Barker JL, Frost JW. Microbial synthesis of p-hydroxybenzoic acid from glucose. Biotechnol Bioeng. 2001;76:376–90.

    Article  CAS  PubMed  Google Scholar 

  62. Bongaerts J, Kramer M, Muller U, Raeven L, Wubbolts M. Metabolic engineering for microbial production of aromatic amino acids and derived compounds. Metab Eng. 2001;3:289–300.

    Article  CAS  PubMed  Google Scholar 

  63. Draths KM, Frost JW. Environmentally compatible synthesis of adipic acid from d-glucose. J Am Chem Soc. 1994;116:399–400.

    Article  CAS  Google Scholar 

  64. Frost JW, Draths KM. Biocatalytic syntheses of aromatics from d-glucose: renewable microbial sources of aromatic compounds. Annu Rev Microbiol. 1995;49:557–79.

    Article  CAS  PubMed  Google Scholar 

  65. Pugh S, McKenna R, Osman M, Thompson B, Nielsen DR. Rational engineering of a novel pathway for producing the aromatic compounds p-hydroxybenzoate, protocatechuate, and catechol in Escherichia coli. Process Biochem. 2014;49:1843–50.

    Article  CAS  Google Scholar 

  66. Snell KD, Draths KM, Frost JW. Synthetic modification of the Escherichia coli chromosome: enhancing the biocatalytic conversion of glucose into aromatic chemicals. J Am Chem Soc. 1996;118:5605–14.

    Article  CAS  Google Scholar 

  67. Li K, Frost JW. Synthesis of vanillin from glucose. J Am Chem Soc. 1998;120:10545–6.

    Article  CAS  Google Scholar 

  68. Kunjapur AM, Tarasova Y, Prather KL. Synthesis and accumulation of aromatic aldehydes in an engineered strain of Escherichia coli. J Am Chem Soc. 2014;136:11644–54.

    Article  CAS  PubMed  Google Scholar 

  69. Balderas-Hernandez VE, Sabido-Ramos A, Silva P, Cabrera-Valladares N, Hernandez-Chavez G, Baez-Viveros JL, Martinez A, Bolivar F, Gosset G. Metabolic engineering for improving anthranilate synthesis from glucose in Escherichia coli. Microb Cell Fact. 2009;8:19.

    Article  PubMed  PubMed Central  Google Scholar 

  70. Albermann C, Ghanegaonkar S, Lemuth K, Vallon T, Reuss M, Armbruster W, Sprenger GA. Biosynthesis of the vitamin E compound δ-tocotrienol in recombinant Escherichia coli cells. ChemBioChem. 2008;9:2524–33.

    Article  CAS  PubMed  Google Scholar 

  71. Kim B, Park H, Na D, Lee SY. Metabolic engineering of Escherichia coli for the production of phenol from glucose. Biotechnol J. 2014;9:621–9.

    Article  CAS  PubMed  Google Scholar 

  72. Magnus J. Method for producing phenol from renewable resources by fermentation. Patent WO 2014076113 A1; 2013.

  73. Miao L, Li Q, Diao A, Zhang X, Ma Y. Construction of a novel phenol synthetic pathway in Escherichia coli through 4-hydroxybenzoate decarboxylation. Appl Microbiol Biotechnol. 2015;99:5163–73.

    Article  CAS  PubMed  Google Scholar 

  74. Thompson B, Machas M, Nielsen DR. Engineering and comparison of non-natural pathways for microbial phenol production. Biotechnol Bioeng. 2016;113:1745–54.

    Article  CAS  PubMed  Google Scholar 

  75. Lin Y, Sun X, Yuan Q, Yan Y. Extending shikimate pathway for the production of muconic acid and its precursor salicylic acid in Escherichia coli. Metab Eng. 2014;23:62–9.

    Article  CAS  PubMed  Google Scholar 

  76. Koma D, Yamanaka H, Moriyoshi K, Ohmoto T, Sakai K. Production of aromatic compounds by metabolically engineered Escherichia coli with an expanded shikimate pathway. Appl Environ Microbiol. 2012;78:6203–16.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Vannelli T, Qi WW, Sweigard J, Gatenby AA, Sariaslani FS. Production of p-hydroxycinnamic acid from glucose in Saccharomyces cerevisiae and Escherichia coli by expression of heterologous genes from plants and fungi. Metab Eng. 2007;9:142–51.

    Article  CAS  PubMed  Google Scholar 

  78. McKenna R, Nielsen DR. Styrene biosynthesis from glucose by engineered E. coli. Metab Eng. 2011;13:544–54.

    Article  CAS  PubMed  Google Scholar 

  79. Noda S, Kondo A. Recent advances in microbial production of aromatic chemicals and derivatives. Trends Biotechnol. 2017;35:785–96.

    Article  CAS  PubMed  Google Scholar 

  80. Vargas-Tah A, Martinez LM, Hernandez-Chavez G, Rocha M, Martinez A, Bolivar F, Gosset G. Production of cinnamic and p-hydroxycinnamic acid from sugar mixtures with engineered Escherichia coli. Microb Cell Fact. 2015;14:6.

    Article  PubMed  PubMed Central  Google Scholar 

  81. Satoh Y, Tajima K, Munekata M, Keasling JD, Lee TS. Engineering of a tyrosol-producing pathway, utilizing simple sugar and the central metabolic tyrosine, in Escherichia coli. J Agric Food Chem. 2012;60:979–84.

    Article  CAS  PubMed  Google Scholar 

  82. Bloch SE, Schmidt-Dannert C. Construction of a chimeric biosynthetic pathway for the de novo biosynthesis of rosmarinic acid in Escherichia coli. ChemBioChem. 2014;15:2393–401.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Jiang M, Zhang H. Engineering the shikimate pathway for biosynthesis of molecules with pharmaceutical activities in E. coli. Curr Opin Biotechnol. 2016;42:1–6.

    Article  PubMed  Google Scholar 

  84. Rodrigues AL, Trachtmann N, Becker J, Lohanatha AF, Blotenberg J, Bolten CJ, Korneli C, Lima AOD, Porto LM, Sprenger GA, Wittmann C. Systems metabolic engineering of Escherichia coli for production of the antitumor drugs violacein and deoxyviolacein. Metab Eng. 2013;20:29–41.

    Article  CAS  PubMed  Google Scholar 

  85. Nakagawa A, Minami H, Kim JS, Koyanagi T, Katayama T, Sato F, Kumagai H. A bacterial platform for fermentative production of plant alkaloids. Nat Commun. 2011;2:326.

    Article  PubMed  Google Scholar 

  86. Nakagawa A, Matsumura E, Koyanagi T, Katayama T, Kawano N, Yoshimatsu K, Yamamoto K, Kumagai H, Sato F, Minami H. Total biosynthesis of opiates by stepwise fermentation using engineered Escherichia coli. Nat Commun. 2016;7:10390.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Wang J, Shen X, Rey J, Yuan Q, Yan Y. Recent advances in microbial production of aromatic natural products and their derivatives. Appl Microbiol Biotechnol. 2018;102:47–61.

    Article  CAS  PubMed  Google Scholar 

  88. Kawaguchi H, Ogino C, Kondo A. Microbial conversion of biomass into bio-based polymers. Bioresour Technol. 2017;245:1664–73.

    Article  CAS  PubMed  Google Scholar 

  89. Sprenger GA. Aromatic amino acids. In: Wendisch VF, editor. Amino acid biosynthesis—pathways, regulation and metabolic engineering. Berlin: Springer; 2007. p. 93–127 (Microbiology monographs).

    Chapter  Google Scholar 

  90. Bailey JE. Toward a science of metabolic engineering. Science. 1991;252:1668–75.

    Article  CAS  PubMed  Google Scholar 

  91. Powell KS, Cooper PD, Forneck A. The biology, physiology and host-plant interactions of grape phylloxera Daktulosphaira vitifoliae. Behav Physiol Root Herbiv. 2013;45:159–218.

    Article  Google Scholar 

  92. Lutke-Eversloh T, Stephanopoulos G. l-Tyrosine production by deregulated strains of Escherichia coli. Appl Microbiol Biotechnol. 2007;75:103–10.

    Article  PubMed  Google Scholar 

  93. Koketsu K, Mitsuhashi S, Tabata K. Identification of homophenylalanine biosynthetic genes from the cyanobacterium Nostoc punctiforme PCC73102 and application to its microbial production by Escherichia coli. Appl Environ Microbiol. 2013;79:2201–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Munoz AJ, Hernandez-Chavez G, de Anda R, Martinez A, Bolivar F, Gosset G. Metabolic engineering of Escherichia coli for improving l-3,4-dihydroxyphenylalanine (l-DOPA) synthesis from glucose. J Ind Microbiol Biotechnol. 2011;38:1845–52.

    Article  CAS  PubMed  Google Scholar 

  95. Wei T, Cheng BY, Liu JZ. Genome engineering Escherichia coli for L-DOPA overproduction from glucose. Sci Rep. 2016;6:30080.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Muller U, van Assema F, Gunsior M, Orf S, Kremer S, Schipper D, Wagemans A, Townsend CA, Sonke T, Bovenberg R, Wubbolts M. Metabolic engineering of the E. coli l-phenylalanine pathway for the production of d-phenylglycine (d-Phg). Metab Eng. 2006;8:196–208.

    Article  PubMed  Google Scholar 

  97. Meiswinkel TM, Rittmann D, Lindner SN, Wendisch VF. Crude glycerol-based production of amino acids and putrescine by Corynebacterium glutamicum. Bioresour Technol. 2013;145:254–8.

    Article  CAS  PubMed  Google Scholar 

  98. Escalante A, Salinas Cervantes A, Gosset G, Bolivar F. Current knowledge of the Escherichia coli phosphoenolpyruvate-carbohydrate phosphotransferase system: peculiarities of regulation and impact on growth and product formation. Appl Microbiol Biotechnol. 2012;94:1483–94.

    Article  CAS  PubMed  Google Scholar 

  99. Herrmann KM. The shikimate pathway as an entry to aromatic secondary metabolism. Plant Physiol. 1995;107:7–12.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Andersen KB, von Meyenburg K. Are growth rates of Escherichia coli in batch cultures limited by respiration? J Bacteriol. 1980;144:114–23.

    CAS  PubMed  PubMed Central  Google Scholar 

  101. Yazdani SS, Gonzalez R. Anaerobic fermentation of glycerol: a path to economic viability for the biofuels industry. Curr Opin Biotechnol. 2007;18:213–9.

    Article  CAS  PubMed  Google Scholar 

  102. Ahn JO, Lee HW, Saha R, Park MS, Jung JK, Lee DY. Exploring the effects of carbon sources on the metabolic capacity for shikimic acid production in Escherichia coli using in silico metabolic predictions. J Microbiol Biotechnol. 2008;18:1773–84.

    CAS  PubMed  Google Scholar 

  103. Salgado H, Gama-Castro S, Peralta-Gil M, Diaz-Peredo E, Sanchez-Solano F, Santos-Zavaleta A, Martinez-Flores I, Jimenez-Jacinto V, Bonavides-Martinez C, Segura-Salazar J, et al. RegulonDB (version 5.0): Escherichia coli K-12 transcriptional regulatory network, operon organization, and growth conditions. Nucleic Acids Res. 2006;34:D394–7.

    Article  CAS  PubMed  Google Scholar 

  104. Sambrook J, Fritsch EF, Maniatis T. Molecular cloning: a laboratory manual. Cold Spring Harbor: Cold Spring Harbor Laboratory Press; 2001.

    Google Scholar 

Download references

Authors’ contributions

BMN, NT and JWY performed the experiments. BMN, GAS and JWY wrote the final manuscript. All authors read and approved the final manuscript.

Acknowledgements

The authors are grateful to Dr. Josef Altenbuchner (Institute of Industrial Genetics, University of Stuttgart) for providing the plasmid pJeM2 and to Dr. Bernd A. Nebel (Institute of Technical Biochemistry, University of Stuttgart) for helping with the MS measurements. We thank Ursula Degner (Research Center Jülich, Germany) for help with the cloning work leading to plasmids pF-109 and pF-112 and Alexander Dietrich (Institute of Biochemical Engineering, University of Stuttgart) for technical assistance with the bioreactor. We thank the Ministerium für Wissenschaft und Kunst Baden-Wuerttemberg for financial support.

Competing interests

The authors declare that they have no competing interests.

Availability of data and materials

The dataset(s) supporting the conclusions of this article are all included within the article and Additional files.

Consent for publication

Not applicable.

Ethics approval and consent to participate

Not applicable.

Funding

Behrouz Mohammadi Nargesi was supported by a personal grant of the Ministry of Science, Research and Technology of Iran (MSRT; 91000277).

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jung-Won Youn.

Additional files

Additional file 1.

 Construction and sequences of vectors

Additional file 2.

Comparison of mass spectra of commercially available L-PAPA (A) with fermentatively produced L-PAPA from E. coli FUS4.7R/pC53BC/pJNTaroFBL (B). The mass spectra were recorded with an Agilent 6130 mass spectrometer system with electron spray ionization in positive mode (Agilent Technologies; Germany). In both measurements the L-PAPA+H+ mass of 181 was detected.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mohammadi Nargesi, B., Trachtmann, N., Sprenger, G.A. et al. Production of p-amino-l-phenylalanine (l-PAPA) from glycerol by metabolic grafting of Escherichia coli. Microb Cell Fact 17, 149 (2018). https://doi.org/10.1186/s12934-018-0996-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12934-018-0996-6

Keywords