You are on page 1of 12

Bioprocess and Biosystems Engineering

https://doi.org/10.1007/s00449-020-02325-5

RESEARCH PAPER

Characterization of two new aromatic amino acid lyases


from actinomycetes for highly efficient production of p‑coumaric acid
Peiwu Cui1,2,3 · Weihong Zhong1 · Yong Qin4 · Fuping Tao4 · Wei Wang2 · Jixun Zhan2,3

Received: 20 December 2019 / Accepted: 4 March 2020


© Springer-Verlag GmbH Germany, part of Springer Nature 2020

Abstract
p-Coumaric acid (p-CA) is a bioactive natural product and an important industrial material for pharmaceuticals and nutraceu-
ticals. It can be synthesized from deamination of l-tyrosine by tyrosine ammonia lyase (TAL). In this work, we discovered
two aromatic amino acid lyase genes, Sas-tal and Sts-tal, from Saccharothrix sp. NRRL B-16348 and Streptomyces sp. NRRL
F-4489, respectively, and expressed them in Escherichia coli BL21(DE3). The two enzymes were functionally characterized
as TAL. The optimum reaction temperature for Sas-TAL and Sts-TAL is 55 °C and 50 °C, respectively; while, the optimum
pH for both TALs is 11. Sas-TAL had a kcat/Km value of 6.2 μM−1 min−1, while Sts-TAL had a much higher efficiency with
a kcat/Km value of 78.3 μM−1 min−1. Both Sts-TAL and Sas-TAL can also take l-phenylalanine as the substrate to yield
trans-cinnamic acid, and Sas-TAL showed much higher phenylalanine ammonia lyase activity than Sts-TAL. Using E. coli/
Sts-TAL as a whole-cell biocatalyst, the productivity of p-CA reached 2.88 ± 0.12 g (L h)−1, which represents the highest
efficiency for microbial production of p-CA. Therefore, this work not only reports the identification of two new TALs from
actinomycetes, but also provides an efficient way to produce the industrially valuable material p-CA.

Keywords Actinomycetes · p-Coumaric acid · Escherichia coli · Phenylalanine ammonia lyase · Tyrosine ammonia lyase

Introduction [1, 2]. However, many of these bioactive natural products


are present in plants at low concentrations, which makes
Phenylpropanoids, one of the largest categories of plant extraction of these compounds from plants inefficient and
secondary metabolites, have received considerable atten- cost-ineffective [3]. Chemical synthesis of these compounds
tion due to their antioxidant, anticancer, anti-atherosclerosis, often involves toxic chemicals and harsh conditions, which
anti-inflammatory and antiviral properties and wide utili- is, thus, not environment-friendly and sustainable. With the
zation in the fields of pharmaceuticals and nutraceuticals development of the biosynthetic technology, microbial pro-
duction of useful phenylpropanoids has become an attractive
approach to producing these valuable products.
* Wei Wang As one of the most important intermediate metabolites in
wangwei402@hotmail.com phenylpropanoid biosynthesis, p-coumaric acid (p-CA) is a
* Jixun Zhan highly significant chemical building block for the produc-
jixun.zhan@usu.edu tion of a variety of valuable substances such as flavonoids,
1
College of Biotechnology and Bioengineering, Zhejiang polyphenols, coumarins, stilbenes, lignans, and polymeric
University of Technology, Hangzhou 310030, Zhejiang, materials [4–9]. Besides, p-CA itself is also a bioactive agent
China exhibiting protective effects against carcinogenesis, athero-
2
TCM and Ethnomedicine Innovation and Development sclerosis, oxidative cardiac damage, neuronal injury and
Laboratory, School of Pharmacy, Hunan University anti-inflammatory activities [10, 11]. p-CA can be formed
of Chinese Medicine, Changsha 410208, Hunan, China directly from non-oxidative deamination of l-tyrosine cat-
3
Department of Biological Engineering, Utah State alyzed by tyrosine ammonia lyase (TAL, Fig. 1) or from
University, 4105 Old Main Hill, Logan, UT 84322‑4105, deamination of l-phenylalanine by phenylalanine ammonia
USA
lyase (PAL, Fig. 1) followed by the hydroxylation catalyzed
4
Hangzhou Viablife Biotech Co., Ltd., 1 Jingyi Road, Yuhang by cinnamate-4-hydroxylse (C4H) [5]. However, expression
District, Hangzhou 311113, Zhejiang, China

13
Vol.:(0123456789)
Bioprocess and Biosystems Engineering

studied. The engineered E. coli strain harboring the Sts-tal


gene was found to be efficient in producing p-CA, represent-
ing a promising strain for the production of this industrially
valuable compound.

Materials and methods

Materials

Phusion High-Fidelity DNA polymerase, restriction


enzymes and T4 DNA ligase were purchased from New Eng-
land Biolabs. l-tyrosine, l-phenylalanine, trans-cinnamic
Fig. 1  Enzymatic synthesis of p-coumaric acid/trans-cinnamic acid acid and p-CA were purchased from Sigma-Aldrich (St.
by TAL and PAL. TAL tyrosine ammonia lyase, PAL phenylalanine Louis, MO, USA). His Pur™ Ni-NTA resin and Luria–Ber-
ammonia lyase, C4H cinnamate-4-hydroxylse tani (LB) medium were purchased from Thermo Fisher Sci-
entific (Rockford, IL, USA). Bradford assay solution was
purchased from TCI America (St. Portland, OR, USA).
of C4H, a cytochrome P450 hydroxylase, in a heterologous Except high-performance liquid chromatography (HPLC)
prokaryotic system is often not satisfactory and its func- solvents, all other chemicals used were of analytical grade
tion requires an efficient electron transfer system [1, 12]. and were purchased from Fisher Scientific (Rockford, IL,
Therefore, the most effective route for p-CA production is USA).
through deamination of l-tyrosine by TAL. Both TAL and
PAL belong to the aromatic amino acid lyase (AAAL) fam- Vectors and strains
ily, which is classified by the substrate specificity as histi-
dine ammonia lyase (HAL, EC 4.3.1.3), TAL (EC 4.3.1.23), The vectors pJET1.2 (Thermo Fisher Scientific, USA) and
PAL (EC 4.3.1.24) or PAL/TAL (EC 4.3.1.25) [13]. Unlike pET28a (Millipore Sigma, USA) were, respectively, used
other AAALs, TAL is relatively rare, especially in bacteria for general cloning and gene expression. E. coli XL-1 Blue
[14]. To our knowledge, TAL has only been identified in and E. coli BL21(DE3) were routinely grown in LB medium
Rhodobacter capsulatus (Integrated Genomics accession at 37 °C and used for plasmid amplification and enzyme
number: RRC01844) [15], Rhodobacter sphaeroides (Gen- expression, respectively. Saccharothrix sp. NRRL B-16348
Bank accession number: YP_355075) [16], Saccharothrix and Streptomyces sp. NRRL F-4489 were obtained from the
espanaensis (GenBank accession number: ABC88669) Agricultural Research Service Culture Collection of USDA,
[17], Rhodotorula glutinis (GenBank accession number: and were grown in YM medium at 28 °C.
M18261.1) [18], Phanerochaete chrysosporium [19], and
Streptomyces sp. (GenBank accession number: JN587500) Amplification of the two putative aromatic amino
[5]. Discovery of more efficient TALs will facilitate engi- acid lyase genes
neered production of p-CA in microbial systems.
In this study, two new actinomycete AAALs were discov- The primers designed for cloning of Sas-tal (1485 bp) and
ered from Saccharothrix sp. NRRL B-16348 (Sas-tal; Gen- Sts-tal (1479 bp) were synthesized by Thermo Fisher Sci-
Bank accession number: WP_053717807.1) and Streptomy- entific and are shown in Table 1. The genomic DNA was
ces sp. NRRL F-4489 (Sts-tal; GenBank accession number: extracted from Saccharothrix sp. NRRL B-16348 and Strep-
WP_066973014.1), respectively. They were heterologously tomyces sp. NRRL F-4489 using the Quick-DNA™ Fungal/
expressed in Escherichia coli BL21 (DE3) and functionally Bacterial Microprep Kit (Zymo Research, USA). To amplify
characterized. Their enzymatic properties on tyrosine were the AAAL genes, the primers (1  μM), genomic DNA

Table 1  Primers for Gene Primers Restriction site


amplification of the two AAAL
genes from Saccharothrix Sas-tal-F 5′-AACAT​ATG​ACC​GAC​GCC​CCC​GTG​GAC​CT-3′ NdeI
sp. NRRL B-16348 and
Sas-tal-R 5′-TAAAG​CTT​CTC​GAG​TCA​GTA​CCG​GCC​CCC​GGT​GACCA-3′ HindIII
Streptomyces sp. NRRL F-4489
Sts-tal-F 5′-AACAT​ATG​CCG​AGT​CTG​GAC​AGCAT-3′ NdeI
Sts-tal-R 5′-TAAAG​CTT​CTC​GAG​TCA​GGC​GGC​GCC​GGT​CAG​CT-3′ HindIII

13
Bioprocess and Biosystems Engineering

(0.2 μL), dNTP mix (200 μM), 5 × buffer (4 μL), DMSO finally eluted with elution buffer (50 mM Tris–HCl, 2 mM
(0.4 μL), Phusion High-Fidelity DNA Polymerase (0.2 μL at EDTA, pH 7.9) containing 100 mM and 50 mM imidazole,
2 U μL−1) and nuclease-free water were mixed to bring the respectively.
volume to 20 μL. The PCR conditions are as follows: initial
denaturation at 98 °C for 5 min, 30 cycles at 98 °C for 40 s,
56 °C for 40 s, and 72 °C for 3.5 min, and finally, 72 °C for In vivo functional characterization of Sas‑TAL
10 min of elongation. and Sts‑TAL

The engineered E. coli cells harboring the Sas-tal or Sts-


Plasmid construction
tal gene were grown and induced as described above. After
induction for 12 h, cells were harvested by centrifugation of
After recovery from a 0.8% agarose gel with a GeneJET
the broth at 3000×g and 4 °C for 10 min and subsequently
Gel Extraction Kit (Thermo Fisher Scientific), the target
washed twice with distilled water. Whole-cell biotransfor-
PCR products were ligated into the pJET1.2 vector to yield
mation was performed using a 10-mL system, which con-
pJET1.2-Sas-tal and pJET1.2-Sts-tal, respectively. After the
sisted of the harvested E. coli cells ­(OD600 1.0) and 3 mM
plasmids were confirmed by digestion check and sequencing,
l-tyrosine in 10 mM sodium phosphate buffer (pH 8.0). The
the two genes were excised from the corresponding plasmids
reaction mixture was incubated at 28 °C and 250 rpm for
using NdeI and HindIII and ligated into pET28a between the
6 h. 3 M HCl (2.0 mL) was then added to terminate the reac-
same sites to construct pET28a-Sas-tal and pET28a-Sts-tal.
tion. The reaction mixture was centrifuged at 13,000×g for
The ligation products were transferred into E. coli XL1-Blue
15 min, and 50 μL of the supernatant was analyzed by HPLC
competent cells through chemical transformation, and the
on an Agilent 1200 HPLC system (Agilent, USA) equipped
transformants were selected on LB agar plate containing
with an Agilent Eclipse Plus ­C18 column (4.6 mm × 250 mm,
50 μg mL−1 kanamycin. The correct plasmids were con-
5 μm) and a diode array detector. The HPLC was eluted with
firmed by digestion check with NdeI and HindIII.
15% (v/v) acetonitrile–water at 30 °C and the product was
detected at 310 nm. The ESI–MS spectrum was collected
Expression and purification of recombinant on an Agilent 6130 single quadruple mass spectrometer. A
enzymes standard curve was established using commercial p-CA as
the standard to quantify the production of p-CA.
The two expression plasmids were, respectively, trans- For in vivo PAL activity assay, the experiment was per-
ferred into E. coli BL21(DE3) competent cells. The result- formed in a similar way except that 3 mM l-phenylalanine
ant strains were subsequently cultivated in LB broth sup- was added as the substrate. HPLC analysis for PAL activ-
plemented with 50  μg  mL −1 kanamycin. The cultures ity was programmed with a gradient of methanol–water
were shaken at 250 rpm and 37 °C until the O ­ D600 reached containing 0.1% formic acid (0–5 min, 5:95; 5–14 min,
0.4–0.6. The cultivation temperature was then decreased to 5:95–20:80; 14–18 min, 20:80–80:20; 18–25 min, 80:20;
28 °C and the cells were induced with 200 μM isopropyl-β- v/v) over 25 min at 278 nm.
d-1-thiogalactopyranoside (IPTG) for 12 h. The cells were
harvested by centrifugation at 3000×g and 4 °C for 10 min,
and subsequently washed twice with distilled water and In vitro functional characterization of recombinant
resuspended in 20 mM pH 7.9 Tris–HCl buffer containing enzymes
0.5 M NaCl. The harvested cells were disrupted by sonica-
tion on ice. The cell lysates were centrifuged at 10,000×g Enzymatic assays were performed in a 200-μL system
and 4 °C for 10 min to collect the soluble fraction contain- containing 10 mM glycine–NaOH buffer (pH 11.0), 3 mM
ing target protein. The supernatants were subsequently sub- l-tyrosine, 6 μM enzyme for purified Sas-TAL and 3 μM
jected to 15% SDS-PAGE analysis using a premixed protein enzyme for purified Sts-TAL. The mixture was incubated at
marker (EZ-Run™ Pre-Stained Rec Protein Ladder, molecu- 50 °C for 1 h before the reaction was terminated by boiling at
lar range: 10–170 kDa, Fisher BioReagents) as the reference. 100 °C for another 10 min. Then, the denatured protein was
Gels were stained with Coomassie Brilliant Blue R250. removed by centrifugation at 13,000×g for 15 min, and 50
To purify the enzymes, the supernatants were loaded μL of the supernatant was analyzed by HPLC for the forma-
onto a HisPur™ Ni-NTA affinity column (Thermo Sci- tion of p-CA with the methods mentioned above. Similarly,
entific, Rockford, USA) according to the manufacturer’s the PAL activity of Sas-TAL and Sts-TAL was determined
protocols. After eluting the column with cool (4 °C) wash- using l-phenylalanine as the substrate. A standard curve was
ing buffer (50 mM Tris–HCl, 2 mM EDTA, 10 mM imida- established using commercial trans-cinnamic acid to meas-
zole, pH 7.9), the recombinant Sas-TAL and Sts-TAL were ure the production of trans-cinnamic acid.

13
Bioprocess and Biosystems Engineering

Determination of the optimal conditions for Sas‑TAL each test was calculated using SPSS 20.0 and indicated
and Sts‑TAL and their kinetic parameters as error bars.

The effects of temperature on the TAL activity of Sas-


TAL and Sts-TAL were determined by measuring the Results and discussion
enzyme activity at various temperatures ranging from 30
to 70 °C under the standard assay conditions in 10 mM Sequence analysis to two putative aromatic amino
glycine–NaOH buffer; while, the effects of pH on the TAL acid lyases from actinomycetes
activity were measured at varying pH values ranging from
7.0 to 12.0 using 10 mM sodium phosphate buffer (pH 7.0 p-CA is a well-known antioxidant from nature. It is also
and 8.0), 10 mM glycine–NaOH buffer (pH 9.0, 10.0, 10.5, a common chemical precursor to synthesize other com-
11.0 and 11.5) and 10 mM potassium chloride-NaOH (pH pounds. TAL-catalyzed deamination of the l -tyrosine
12.0) at 50 °C. is a convenient biological process for the production of
For the kinetic studies, the initial reaction rates of the p-CA. Therefore, discovery of efficient TALs is critical
purified TAL were evaluated by reacting the enzyme with for developing cost-effective production processes of this
different concentrations of l-tyrosine ranging from 100 compound. TALs have been previously found from a few
to 3 mM for both Sas-TAL and Sts-TAL, in 10 mM gly- sources such as R. capsulatus [15] and S. espanaensis [17].
cine–NaOH buffer (pH 11.0) at 50  °C for 30  min. The They are often involved in the biosynthesis of proteins
respective kinetic parameters, including Km and Vmax, were and secondary metabolites. For example, a TAL from R.
calculated using the Lineweaver–Burk plots [20]. The value capsulatus was found to be involved in the biosynthesis
of the turnover number (kcat) was calculated from the equa- of a photoactive yellow protein [15]. Sam8 is a TAL from
tion: kcat = Vmax/[E], where [E] is the enzyme concentration S. espanaensis that is involved in the biosynthesis of caf-
in the reaction system and Vmax is the maximum velocity. feic acid, a part of the aglycon of saccharomicins A and
Furthermore, the effects of metal ions on the activity of B [17]. Discovery and heterologous expression of these
TAL were also analyzed by testing the purified enzymes TALs makes it possible to produce p-CA in engineered
with 3  mM salt solutions of NaCl, KCl, F ­ eSO4, ­FeCl3, microorganisms such as bacteria and yeast [13].
­ZnCl2, ­CaCl2, ­CuCl2, ­MgCl2 and M ­ nCl2. The TAL activity Genome mining of potential TALs led to the discovery
was studied under standard reaction conditions with 10 mM of two putative AAAL genes from two actinomycetes.
glycine–NaOH buffer (pH 11.0), 3 mM l-tyrosine, 6 μM The 1485 bp Sas-tal gene was found in Saccharothrix sp.
enzyme for purified Sas-TAL and 3 μM enzyme for purified NRRL B-16348 and the 1479 bp Sts-tal gene was found
Sts-TAL at 50 °C. from Streptomyces sp. NRRL F-4489. BLAST analysis of
the protein sequences of these two proteins revealed that
Production of p‑CA from l‑tyrosine using they are homologous to a number of TALs and unchar-
engineered E. coli cells as the whole‑cell biocatalysts acterized AAALs. Phylogenetic analysis of Sas-TAL and
Sts-TAL with 11 other proteins including TALs, PALs and
The effects of temperature and pH on the efficiency of TALs/PALs from different organisms (Fig. 2a) showed
whole-cell biotransformation were also studied with a simi- that Sas-TAL and Sts-TAL belong to two different but
lar method mentioned above. The tested temperatures ranged closely related clades. Both of them showed closer rela-
from 30 to 70 °C and the pH values from 7.0 to 12.0. tionship to reported TALs rather than PALs. Multiple pro-
Two different cell concentrations and substrate concentra- tein sequence alignment (Fig. 2b) revealed the presence
tions were also tested. The cells (­ OD600 1.0 or 10.0) were of the conserved Ala-Ser-Gly motif that forms the unique
reacted with l-tyrosine (3 mM or 30 mM) in 10 mM gly- modified amino acid cofactor 3,5-dihyro-5-methylidene-
cine–NaOH buffer (pH 11.0) at 50 °C for 1 h, with a total 4H-imidazol-4-one (MIO) [14]. It was reported that there
volume of 10.0 mL. The bioconversion was terminated by is a substrate selectivity switch region (boxed in Fig. 2b)
addition of 2.0 mL of 3 M HCl into the system. The reaction in AAALs. In general, PALs have FL (phenylalanine and
mixtures were centrifuged at 13,000×g for 15 min and the leucine) in this region, TALs have HL (histidine and leu-
supernatants were subjected to HPLC analysis. cine), and PALs/TALs have HQ (histidine and glutamine),
respectively [21]. A previous work reported that replac-
Statistical analysis ing the active site His residue of TAL with Phe switched
its substrate specify from l-tyrosine to l-phenylalanine.
All experiments were carried out in triplicate and the Similarly, mutation of P ­ he 144 with His in Arabidopsis
mean values were calculated. The standard deviation for thaliana PAL1 converted it into a selective TAL [14].

13
Bioprocess and Biosystems Engineering

Fig. 2  Phylogenetic and sequence analysis of Sas-TAL and Sts-TAL. [26]; 9, Tp-PAL from Trifolium pretense (AAZ29733.1) [27]; 10,
a Phylogenetic analysis of Sas-TAL, Sts-TAL and 11 other aromatic Ib-PAL from Inonotus baumii (OCB91043.1) [28]; 11, Rg-PAL
amino acid lyases (AAALs). b Partial amino acid sequence align- from Rhodotorula glutinis (AHB63479.1) [29]; 12, Zm-PAL from
ment of Sas-TAL, Sts-TAL and 11 other AAALs. The enzymes and Zea mays L. cv Corso (AAL40137.1) [30]; 13, Tc-PAL from Tri-
GenBank accession numbers are: 1, Sas-TAL from Saccharothrix sp. chosporon cutaneum (ABA69898.1) [21]. Sas: Saccharothrix sp.;
NRRL B-16348 (WP_053717807.1); 2, Sts-TAL from Streptomyces Sts: Streptomyces sp.; Rc: Rhodobacter capsulatus; Rs: Rhodobac-
sp. NRRL F-4489 (WP_066973014.1); 3, Rc-TAL from Rhodobacter ter sphaeroides; Se: Saccharothrix espanaensis; Pc: Petroselinum
capsulatus (WP_013066811.1) [15]; 4, Rs-TAL from Rhodobacter crispum; Jr: Juglans regia; Jc: Jatropha curcas; Tp: Trifolium pre-
sphaeroides (WP_011339422.1) [16]; 5, Se-TAL from Saccharothrix tense; Ib: Inonotus baumii; Rg: Rhodotorula glutinis; Zm: Zea mays;
espanaensis (WP_015103237.1) [11, 17]; 6, Pc-PAL from Petroseli- Tc: Trichosporon cutaneum. The substrate selectivity switch is boxed
num crispum Nym (CAA57056.1) [24]; 7, Jr-PAL from Juglans regia and the MIO-forming Ala-Ser-Gly motif is asterisked
(JX069977.1) [25]; 8, Jc-PAL from Jatropha curcas L. (ABI33979.1)

Therefore, the active site His residue plays a critical role analysis and protein sequence alignment, Sas-TAL and
in the TAL activity and substrate selectivity. Both Sas- Sts-TAL were predicted to be TALs that catalyze deami-
TAL and Sts-TAL have HL in this region, suggesting that nation of l-tyrosine to yield p-CA.
they are TALs. Based on the results of the phylogenetic

13
Bioprocess and Biosystems Engineering

and pET28a-Sts-tal. The two plasmids were expressed in


E. coli BL21(DE3) with IPTG induction. The cells were
lysed by sonication for SDS-PAGE analysis. As shown in
Fig. 3a, compared to the vector control (E. coli BL21(DE3)/
pET28a, lane 1), a new ~ 55 kb protein band showed up in
the lysate of E. coli BL21(DE3)/pET28a-Sas-tal (lane 2),
suggesting that SAS-TAL has been successfully expressed.
Similarly, Sts-TAL was also expressed in E. coli BL21(DE3)
(Fig. 3b). These two enzymes were purified by Ni-NTA
chromatography to homogeneity (line 3 of Fig. 3a, b). The
yields of Sas-TAL and Sts-TAL were 92.4 ± 0.8 mg L−1 and
Fig. 3  SDS-PAGE analysis of the expression and purification of Sas- 202.2 ± 5.7 mg L−1, respectively.
TAL and Sts-TAL. a Expression and purification of Sas-TAL from
E. coli BL21(DE3)/pET28a-Sas-tal. b Expression and purification
of Sas-TAL from E. coli BL21(DE3)/pET28a-Sts-tal. M: protein In vivo and in vitro functional characterization
ladder; 1: lysate of E. coli BL21(DE3)/pET28a; 2. lysate of E. coli of Sas‑TAL and Sts‑TAL
BL21(DE3)/pET28a-Sas-tal (a) or -Sts-tal (b); 3: purified Sas-TAL
(a) or Sts-TAL (b)
To verify the function of recombinant TALs, 3 mM l-tyros-
ine was, respectively, reacted with IPTG-induced E. coli
Expression and purification of Sas‑TAL and Sts‑TAL BL21(DE3)/pET28a-Sas-tal, E. coli BL21(DE3)/pET28a-
Sts-tal and E. coli BL21(DE3)/pET28a in 10 mM sodium
The two genes were amplified from the genomic DNA phosphate buffer (pH 8.0). HPLC analysis showed that the
of corresponding hosts, and ligated to pET28a to yield vector control did not convert the substrate into any detect-
the corresponding expression plasmids, pET28a-Sas-tal able product (trace i, Fig. 4a); while, both the engineered

Fig. 4  In vivo functional characterization of Sas-TAL and Sts-TAL. a Sas-tal; (iii) l-tyrosine + E. coli BL21(DE3)/pET28a-Sts-tal; (iv)
HPLC analysis of conversion of l-tyrosine into p-coumaric acid by standard of p-coumaric acid. b UV spectrum of the bioconversion
Sas-TAL and Sts-TAL in E. coli at 310  nm. (i) l-tyrosine + E. coli product at 15.0  min. c ESI–MS (−) spectrum of the bioconversion
BL21(DE3)/pET28a; (ii) l-tyrosine + E. coli BL21(DE3)/pET28a- product at 15.0 min

13
Bioprocess and Biosystems Engineering

strains yielded a less polar product at 15 min (traces ii and be 50 °C. These are higher than the optimum temperatures
iii, Fig. 4a). This product has the same retention time as the of several reported TALs that are typically in the range of
commercial standard of p-CA (trace iv, Fig. 4a). Further- 25–40 °C (Table 2).
more, the product showed the same UV spectrum (Fig. 4b) We next tested the effect of pH on the activity of Sas-
as that reported for p-CA [22]. Finally, the ESI–MS spec- TAL and Sts-TAL. Eight pH values, ranging from 7 to
trum (Fig. 4c) of this compound showed a [M−H]− ion peak 12, were examined for the reaction systems of these two
at m/z 163.1, indicating that its molecular weight is 164 Da, enzymes. As shown in Fig.  6b, the enzymatic activ-
which is consistent with that of p-CA. Therefore, this prod- ity of Sts-TAL increased when the pH was increased
uct was characterized as p-CA, and both Sas-TAL and Sts- from 7 to 11, and the maximum activity of Sts-TAL
TAL can catalyze the deamination of l-tyrosine to yield p- (330.57 ± 3.05  µM p-CA ­m g −1  min −1) was achieved at
CA when they were expressed in E. coli BL21(DE3). The pH 11. Apparent drop in the activity of Sts-TAL was
purified enzymes were also reacted with l-tyrosine in vitro observed when the pH was further increased to 11.5 and
and showed the same product, further confirming that these 12. As such, the optimum pH value for Sts-TAL was
two enzymes are TALs. determined to be 11. For Sas-TAL, its enzymatic activity
increased from 11.94 ± 0.97 to 147.99 ± 1.59 µM p-CA
Determination of optimum reaction temperature ­mg−1 min−1 when the pH was increased from 7 to 10.5.
and pH for Sas‑TAL and Sts‑TAL Similar enzymatic activity was observed for Sas-TAL 10.5,
11.0 and 11.5, and a significant decrease in the activity
To determine the kinetic parameters of Sas-TAL and Sts- was observed when the reaction pH was increased to 12.
TAL, we first measured how the reaction temperature and Thus, we chose pH 11 for both enzymes for the following
pH affect the enzymatic activity of these two TALs. A total studies. Alkaline pH is generally preferred by TALs to
of seven temperatures, including 30, 40, 45, 50, 55, 60 and catalyze the deamination reaction of l-tyrosine. The opti-
70 °C, were tested. We established the standard curves of mum pH for Sas-TAL and Sts-TAL is much higher than the
p-CA (Fig. 5a) and cinnamic acid (Fig. 5b) using authen- optimum pH values for several previously reported TALs
tic samples. As shown in Fig. 6a, the activity of Sas-TAL (Table 2), such as the TAL from R. glutinis (pH 8.5) and
increased from 55.50 ± 0.14 to 181.81 ± 8.86 µM p-CA Sam8 from S. espanaensis (pH 8.8). This indicated that
­mg−1 min−1 when the reaction temperature was increased Sas-TAL and Sts-TAL might be more stable and perform
from 30 to 55 °C. However, further increase in the tem- more efficiently at pH 11 than others. We also calculated
perature resulted in decreased enzymatic activity, which the theoretical isoelectric point (pI) values of related TALs
might be due to the inactivation of the enzyme at high tem- using the ExPASy-ProtParam tool. As shown in Table 2,
peratures. Therefore, 55 °C is the optimum temperature both Sas-TAL and Sts-TAL have a much lower pI value
for Sas-TAL for in vitro enzymatic reactions. Similarly, than Rg-TAL, Rc-TAL and Rs-TAL. By contrast, Sam8 has
the optimum temperature for Sts-TAL was determined to a similar pI value with Sas-TAL and Sts-TAL.

Fig. 5  The standard curves of p-CA (a) and trans-cinnamic acid (b) for quantification of the product formation

13
Bioprocess and Biosystems Engineering

Fig. 6  Effects of the temperature (a), pH (b), substrate concentration (c) and metal ions (d) on the activity/reaction velocity of purified Sas-TAL
and Sts-TAL

Table 2  Comparison of optimum temperature and pH as well as kinetic parameters of Sas-TAL and Sts-TAL with other reported TALs

Enzyme Source Optimum Optimum pH Theoretical pI Km (μM) kcat ­(s−1) kcat/Km (μM−1 s−1) References
temperature
(°C)

Sas-TAL Saccharothrix sp. NRRL 55 11.0 5.14 1492.2 155.2 0.104 This work
B-16348
Sts-TAL Streptomyces sp. NRRL F-4489 50 11.0 4.95 336.5 439.1 1.3 This work
Rg-TAL Rhodotorula glutinis 40 8.5 6.44 380.0 114 0.3 [1]
Rc-TAL Rhodobacter capsulatus DSMZ 35 8.5 6.50 15.6 27.7 1.77 [15]
1710
Rs-TAL Rhodobacter sphaeroides 25 8.5 7.30 60 0.02 0.333 × 10–3 [16]
Sam8 Saccharothrix espanaensis 30 8.8 5.27 15.5 0.015 0.968 × 10–3 [17]
NRRL-15764

13
Bioprocess and Biosystems Engineering

Determination of the kinetic parameters of Sas‑TAL TALs shown in Table 2, both Sas-TAL and Sts-TAL exhib-
and Sts‑TAL ited a much higher turnover number (Kcat) toward l-tyros-
ine than reported TALs, and Sts-TAL also showed higher
We next measured the kinetic parameters of Sas-TAL and catalytic efficiency (kcat/Km), indicating that these enzymes,
Sts-TAL at the optimal temperature and pH. Specifically, pH especially Sts-TAL, represent efficient biocatalysts for p-CA
11 and 55 °C were used for Sas-TAL, and pH 11 and 50 °C production.
for Sts-TAL. Figure 6c showed the kinetic behaviors of the
two TALs. The two linear double-reciprocal plots obtained
with higher correlation coefficients indicated that both TALs Effect of metal ions on the enzymatic activity
obeyed Michaelis–Menten kinetics with regard to l-tyrosine. of Sas‑TAL and Sts‑TAL
Based on the nonlinear regression analysis, Sas-TAL showed
Kcat, Km and kcat/Km values of 9.3 × 103 min−1 (or 155.2 s−1), To analyze the effect of metal ions on the activity of Sas-
1492.2 μM and 6.2 μM−1 min−1 (or 0.104 μM−1 s−1), while TAL and Sts-TAL, 3 mM of different metal ions was added
these values for Sts-TAL were 2.6 × 104 min−1 (or 439.1 s−1), to the reaction system. As shown in Fig. 6d, addition of
336.5 μM and 78.3 μM−1 min−1 (or 1.3 μM−1 s−1), respec- ­Na+, ­K+, ­Ca2+ and ­Mg2+ slightly increased the enzymatic
tively. The turnover numbers of Sas-TAL and Sts-TAL are activity of Sas-TAL (less than 10%), while ­Fe2+, ­Fe3+, ­Zn2+,
higher than those of known TALs, and their Km values are ­Cu2+ and M ­ n2+ showed strong inhibition to this enzyme.
larger than most reported TALs as well. Both high kcat and However, there were no ions that increased the activity of
Km values are likely due to the higher reaction temperatures. Sts-TAL. All tested groups showed weaker activity than the
Compared with the kinetic parameters of reported microbial control (showing 25.2–92.2% of the original activity), except

Fig. 7  The PAL activity of Sas-TAL and Sts-TAL. a HPLC analysis Sas-tal; (iii) l-tyrosine + E. coli BL21(DE3)/pET28a-Sts-tal; (iv)
(278  nm) of the formation of cinnamic acid from l-phenylalanine standard of p-coumaric acid. b Comparison of in vitro TAL activity
by Sas-TAL- and Sts-TAL-harboring E. coli. (i) l-tyrosine + E. coli and PAL activity for Sas-TAL and Sts-TAL
BL21(DE3)/pET28a; (ii) l-tyrosine + E. coli BL21(DE3)/pET28a-

13
Bioprocess and Biosystems Engineering

­ a+ and K
N ­ + that had no effect on the activity of Sts-TAL at with 3 mM l-tyrosine in 10 mM glycine–NaOH buffer (pH
3 mM. 11) at different temperatures ranging from 30 to 70 °C for
1 h. As shown in Fig. 8a, when the reaction temperature was
Deamination of l‑phenylalanine by Sas‑TAL increased, enhanced specific activity of E. coli BL21(DE3)/
and Sts‑TAL Sts-TAL was also obtained. At 50 °C, the specific activity
reached 2291.0 ± 17.2 μM p-CA per ­OD600 unit cells h−1.
Both TALs and PALs belong to the family of aro- Further increase in the reaction temperature to 60 °C and
matic amino acid lyases. There is also one type of PAL/TAL 70 °C led to a decrease in the activity. A similar trend was
capable of catalyzing the deamination of both phenylalanine observed for E. coli BL21(DE3)/Sas-TAL, except that its
and tyrosine [13]. We, therefore, investigated whether Sas- overall efficiency was much lower than E. coli BL21(DE3)/
TAL and Sts-TAL can convert l-phenylalanine into trans- Sts-TAL. Therefore, our results suggested that the optimum
cinnamic acid. HPLC analysis showed that compared to the temperature for the production of p-CA by both strains are
control (trace i, Fig. 7a), the whole-cell biotransformation of 50 °C (Fig. 8a). We also evaluated what pH value is opti-
l-phenylalanine with E. coli BL21(DE3)/Sas-TAL generated mal for whole-cell biotransformation of l-tyrosine to p-CA.
a less polar product (trace ii, Fig. 7a). Similarly, the same Eight different pH values (pH 7, 8, 9, 10, 10.5, 11, 11.5 and
product was obtained from the reaction of l-phenylalanine 12) were tested, among which pH 11 was the best for both E.
with E. coli BL21(DE3)/Sas-TAL (trace iii, Fig. 7a). This coli BL21(DE3)/Sas-TAL and E. coli BL21(DE3)/Sts-TAL
product was found to be trans-cinnamic acid by a compari- (Fig. 8b). This is consistent with the optimum pH for the
son with the commercial standard (trace iv, Fig. 7a). These purified enzymes.
results indicated that the two actinomycete TALs also have We next compared two different cell concentrations in the
the PAL activity. whole-cell biotransformation process. E. coli BL21(DE3)/
To compare the TAL and PAL activities of Sas-TAL, Sas-TAL and E. coli BL21(DE3)/Sts-TAL were tested for
in vitro reactions were conducted at pH 11 and 55 °C, and converting l-tyrosine into p-CA at different O ­ D600 values
calculated according to the standard curves (Fig. 5a, b). As ­(OD600 1.0 and 10.0). The system with O ­ D600 1.0 cells used
shown in Fig. 7b, under the same reaction conditions, Sas- 3 mM l-tyrosine as the substrate, while 30 mM l-tyrosine
TAL showed a PAL activity of 169.24 ± 1.71 µM trans-cin- was used for O­ D600 10.0 cells. The reactions were performed
namic acid·mg−1·min−1, which is 36% higher than its TAL at pH 11 and 50 °C for 1 h. As shown in Fig. 8c, when
activity. By contrast, when TAL and PAL activities of Sts- the ­OD600 1.0 cells reacted with 3 mM l-tyrosine, E. coli
TAL were measured at pH 11 and 55 °C, the TAL activity BL21(DE3)/Sas-TAL achieved 0.135 ± 0.001  mM p-CA
was found to be 301.05 ± 4.15 µM p-CA ­mg−1 min−1, and with a 4.5% conversion rate; whereas, E. coli BL21(DE3)/
its PAL activity was only 19.67 ± 0.44 µM trans-cinnamic Sts-TAL achieved 2.538 ± 0.103 mM p-CA with an 84.6%
acid ­mg−1 min−1. The TAL activity of Sts-TAL is 15.3- conversion rate. When the ­OD600 value was increased to 10.0
fold higher than its PAL activity. This results indicate that and l-tyrosine concentration to 30 mM, the final titers of p-
both Sas-TAL and Sts-TAL can catalyze the deamination CA for E. coli BL21(DE3)/Sas-TAL and E. coli BL21(DE3)/
of l-tyrosine and l-phenylalanine. Sts-TAL strongly prefers Sts-TAL were 1.206 ± 0.019 mM (equivalent to 0.198 g L−1)
l-tyrosine as the substrate; whereas, Sas-TAL slightly prefers and 17.575 ± 0.727 mM (equivalent to 2.88 g L−1), with
l-phenylalanine to l-tyrosine. the corresponding conversation rates of 4.0% and 58.6%,
respectively.
Determination of the optimal conditions Consistent with the in vitro enzymatic activities, E. coli
for the production of p‑CA using E. coli BL21(DE3)/ BL21(DE3)/Sts-TAL is much more efficient than E. coli
Sas‑TAL and E. coli BL21(DE3)/Sts‑TAL cells BL21(DE3)/Sas-TAL. This is expected because Sts-TAL
has better catalytic efficiency and expression level than
Since purification of enzymes from E. coli cells is costly, Sas-TAL. The productivity of E. coli BL21(DE3)/Sts-TAL
whole-cell biocatalysis represents a convenient and cost- reached 2.88 ± 0.12 g (L h)−1 (equivalent to 1.76 mM p-CA/
effective way to produce industrially valuable products unit of OD600 cells h−1). This is four times the productivity
[23]. We already showed that the TAL-harboring E. coli of p-CA (440 μM p-CA/unit of OD600 cells) in E. coli by
cells can convert l-tyrosine into p-CA in 3.3. To find out the TALs from Herpetosiphon aurantiacus and Flavobac-
the optimal conditions for the whole-cell biotransforma- terium johnsoniae, which was measured in the induced fer-
tion process, we next investigated the effects of the reac- mentation broth for more than 24 h [13]. Therefore, E. coli
tion temperature and pH on the production of p-CA. To BL21(DE3)/Sts-TAL outperformed the previously reported
test the optimal temperature for whole-cell conversion of bacterial strains for the production of p-CA, representing a
l-tyrosine into p-CA, E. coli BL21(DE3)/Sas-TAL and E. promising strain for the production of this valuable chemi-
coli BL21(DE3)/Sts-TAL cells ­(OD600 1.0) were incubated cal. The production efficiency may be further increased by

13
Bioprocess and Biosystems Engineering

Fig. 8  Determination of optimal conditions for whole-cell conver- of the reaction pH on the specific conversion activity of E. coli
sion of l-tyrosine to p-CA with engineered E. coli cells. a Effect of BL21(DE3)/Sas-TAL and E. coli BL21(DE3)/Sts-TAL. c Compari-
the reaction temperature on the specific conversion activity of E. son of the specific activity of E. coli BL21(DE3)/Sas-TAL and E. coli
coli BL21(DE3)/Sas-TAL and E. coli BL21(DE3)/Sts-TAL. b Effect BL21(DE3)/Sts-TAL at different cell and substrate concentrations

directed evolution of Sts-TAL and enhanced expression of coli strain harboring Sts-tal shows a high conversion rate of
this enzyme. 1.758 mM p-CA/unit of O ­ D600 cells, representing the high-
est productivity of p-CA in prokaryotic cells. This discovery
will provide a much more effective way for producing the
Conclusions industrially valuable material p-CA through an environmen-
tal friendly and sustainable process.
In this work, we identified two AAALs from two different
actinomycetes. Both Sas-TAL and Sts-TAL were found to Acknowledgements  This research was financially supported by
National Natural Science Foundation of China (81973211) and Fund-
have the TAL and PAL activities. While Sas-TAL showed a ing of Changsha Science and Technology Bureau (kq1701069).
better PAL activity than TAL activity, Sts-TAL strongly pre-
fers l-tyrosine as the deamination substrate. The optimum Compliance with ethical standards 
temperatures of Sas-TAL and Sts-TAL are 55 °C and 50 °C,
while the optimum pH for both TALs is 11.0, respectively. Conflict of interest  The authors declare no conflict of interest.
Kinetic studies on the purified enzymes revealed that Sts-
TAL is a much more efficient TAL than Sas-TAL, with a Ethical approval  This study does not contain any studies with human
participants or animal performed by any of the authors.
kcat/Km value of 78.3 μM−1 min−1. Besides, the engineered E.

13
Bioprocess and Biosystems Engineering

References 16. Xue Z, McCluskey M, Cantera K, Sariaslani FS, Huang L (2007)


Identification, characterization and functional expression of a
tyrosine ammonia-lyase and its mutants from the photosynthetic
1. Zhou S, Liu P, Chen J, Du G, Li H, Zhou J (2016) Characteriza-
bacterium Rhodobacter sphaeroides. J Ind Microbiol Biotechnol
tion of mutants of a tyrosine ammonia-lyase from Rhodotorula
34:599–604
glutinis. Appl Microbiol Biotechnol 100:10443–10452
17. Berner M, Krug D, Bihlmaier C, Vente A, Muller R, Bechthold
2. Darsandhari S, Dhakal D, Shrestha B, Parajuli P, Seo JH, Kim TS,
A (2006) Genes and enzymes involved in caffeic acid biosynthe-
Sohng JK (0041BP) Characterization of regioselective flavonoid
sis in the actinomycete Saccharothrix espanaensis. J Bacteriol
O-methyltransferase from the Streptomyces sp. KCTC 0041BP.
188:2666–2673
Enzyme Microb Technol 113:29–36
18. Vannelli T, Wei Qi W, Sweigard J, Gatenby AA, Sariaslani FS
3. Vargas-Tah A, Martinez LM, Hernandez-Chavez G, Rocha M,
(2007) Production of p-hydroxycinnamic acid from glucose in
Martinez A, Bolivar F, Gosset G (2015) Production of cinnamic
Saccharomyces cerevisiae and Escherichia coli by expression of
and p-hydroxycinnamic acid from sugar mixtures with engineered
heterologous genes from plants and fungi. Metab Eng 9:142–151
Escherichia coli. Microb Cell Fact 14:6
19. Xue Z, McCluskey M, Cantera K, Ben-Bassat A, Sariaslani FS,
4. Shin SY, Jung SM, Kim MD, Han NS, Seo JH (2012) Production
Huang L (2007) Improved production of p-hydroxycinnamic acid
of resveratrol from tyrosine in metabolically engineered Saccha-
from tyrosine using a novel thermostable phenylalanine/tyrosine
romyces cerevisiae. Enzyme Microb Technol 51:211–216
ammonia lyase enzyme. Enzyme Microb Technol 42:58–64
5. Zhu Y, Liao S, Ye J, Zhang H (2012) Cloning and characteriza-
20. Hans L, Dean B (1934) The determination of enzyme dissociation
tion of a novel tyrosine ammonia lyase-encoding gene involved
constants. J Am Chem Soc 56:658–666
in bagremycins biosynthesis in Streptomyces sp. Biotechnol Lett
21. Goldson-Barnaby A, Scaman CH (2013) Purification and char-
34:269–274
acterization of phenylalanine ammonia lyase from Trichosporon
6. Rodriguez A, Chen Y, Khoomrung S, Ozdemir E, Borodina I,
cutaneum. Enzyme Res 2013:670702
Nielsen J (2017) Comparison of the metabolic response to over-
22. Appert C, Logemann E, Hahlbrock K, Schmid J, Amrhein N
production of p-coumaric acid in two yeast strains. Metab Eng
(1994) Structural and catalytic properties of the four phenyla-
44:265–272
lanine ammonia-lyase isoenzymes from parsley (Petroselinum
7. Schroeder AC, Kumaran S, Hicks LM, Cahoon RE, Halls C, Yu
crispum Nym.). Eur J Biochem 225:491–499
O, Jez JM (2008) Contributions of conserved serine and tyros-
23. Xu F, Deng G, Cheng S, Zhang W, Huang X, Li L, Cheng H, Rong
ine residues to catalysis, ligand binding, and cofactor process-
X, Li J (2012) Molecular cloning, characterization and expres-
ing in the active site of tyrosine ammonia lyase. Phytochemistry
sion of the phenylalanine ammonia-lyase gene from Juglans regia.
69:1496–1506
Molecules 17:7810–7823
8. Noda S, Kawai Y, Tanaka T, Kondo A (2015) 4-Vinylphenol bio-
24. Gao J, Zhang S, Cai F, Zheng X, Lin N, Qin X, Ou Y, Gu X, Zhu
synthesis from cellulose as the sole carbon source using phenolic
X, Xu Y, Chen F (2012) Characterization, and expression profile
acid decarboxylase- and tyrosine ammonia lyase-expressing Strep-
of a phenylalanine ammonia lyase gene from Jatropha curcas L.
tomyces lividans. Bioresour Technol 180:59–65
Mol Biol Rep 39:3443–3452
9. Fujiwara R, Noda S, Kawai Y, Tanaka T, Kondo A (2016) 4-Vinyl-
25. Wang S, Zhang S, Zhou T, Zeng J, Zhan J (2013) Design and
phenol production from glucose using recombinant Streptomyces
application of an in vivo reporter assay for phenylalanine ammo-
mobaraense expressing a tyrosine ammonia lyase from Rhodobac-
nia-lyase. Appl Microbiol Biotechnol 97:7877–7885
ter sphaeroides. Biotechnol Lett 38:1543–1549
26. Lin W, Liu A, Weng C, Li H, Sun S, Song A, Zhu H (2018) Clon-
10. Kheiry M, Dianat M, Badavi M, Mard SA, Bayati V (2019)
ing and characterization of a novel phenylalanine ammonia-lyase
p-Coumaric acid attenuates lipopolysaccharide-induced lung
gene from Inonotus baumii. Enzyme Microb Technol 112:52–58
inflammation in rats by scavenging ROS production: an in vivo
27. Zhu L, Cui W, Fang Y, Liu Y, Gao X, Zhou Z (2013) Cloning,
and in vitro study. Inflammation 42:1939–1950
expression and characterization of phenylalanine ammonia-lyase
11. Xue Y, Zhang Y, Cheng D, Daddy S, He Q (2014) Genetically
from Rhodotorula glutinis. Biotechnol Lett 35:751–756
engineering Synechocystis sp. Pasteur Culture Collection 6803
28. Rosler J, Krekel F, Amrhein N, Schmid J (1997) Maize phenyla-
for the sustainable production of the plant secondary metabolite
lanine ammonia-lyase has tyrosine ammonia-lyase activity. Plant
p-coumaric acid. Proc Natl Acad Sci USA 111:9449–9454
Physiol 113:175–179
12. Jones JA, Collins SM, Vernacchio VR, Lachance DM, Koffas MA
29. Wang J, Yue YD, Tang F, Sun J (2012) Screening and analysis
(2016) Optimization of naringenin and p-coumaric acid hydroxy-
of the potential bioactive components in rabbit plasma after oral
lation using the native E. coli hydroxylase complex. HpaBC Bio-
administration of hot-water extracts from leaves of Bambusa tex-
technol Prog 32:21–25
tilis McClure. Molecules 17:8872–8885
13. Jendresen CB, Stahlhut SG, Li M, Gaspar P, Siedler S, Forster
30. Yuan H, Wang H, Fidan O, Qin Y, Xiao G, Zhan J (2019) Identi-
J, Maury J, Borodina I, Nielsen AT (2015) Highly Active and
fication of new glutamate decarboxylases from Streptomyces for
specific tyrosine ammonia-lyases from diverse origins enable
efficient production of gamma-aminobutyric acid in engineered
enhanced production of aromatic compounds in bacteria and Sac-
Escherichia coli. J Biol Eng 13:24
charomyces cerevisiae. Appl Environ Microbiol 81:4458–4476
14. Watts KT, Mijts BN, Lee PC, Manning AJ, Schmidt-Dannert
Publisher’s Note Springer Nature remains neutral with regard to
C (2006) Discovery of a substrate selectivity switch in tyrosine
jurisdictional claims in published maps and institutional affiliations.
ammonia-lyase, a member of the aromatic amino acid lyase fam-
ily. Chem Biol 13:1317–1326
15. Kyndt JA, Meyer TE, Cusanovich MA, Van Beeumen JJ (2002)
Characterization of a bacterial tyrosine ammonia lyase, a bio-
synthetic enzyme for the photoactive yellow protein. FEBS Lett
512:240–244

13

You might also like