Establishing an Artificial Pathway for Efficient Biosynthesis of

Dec 28, 2017 - Hydroxytyrosol (HT) is a valuable natural phenolic compound with strong antioxidant activity and various physiological and pharmaceutic...
0 downloads 12 Views 1MB Size
Subscriber access provided by Grand Valley State | University

Article

Establishing an artificial pathway for efficient biosynthesis of hydroxytyrosol Xianglai Li, Zhenya Chen, Yifei Wu, Yajun Yan, Xinxiao Sun, and Qipeng Yuan ACS Synth. Biol., Just Accepted Manuscript • DOI: 10.1021/acssynbio.7b00385 • Publication Date (Web): 28 Dec 2017 Downloaded from http://pubs.acs.org on January 1, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Synthetic Biology is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Synthetic Biology

1

Establishing an artificial pathway for efficient biosynthesis of hydroxytyrosol

2

Xianglai Lia, Zhenya Chena, Yifei Wua, Yajun Yanb, Xinxiao Suna*, Qipeng Yuana*

3

a

State Key Laboratory of Chemical Resource Engineering, Beijing University of Chemical Technology, Beijing 100029, China

4 5

b

College of Engineering, The University of Georgia, Athens, GA 30602, USA

6 7

Corresponding authors:

8

*Xinxiao Sun

9

[email protected]

10

Telephone: +86-10-64431557

11

Address: 15 Beisanhuan East Road, Chaoyang District, Beijing 100029, China

12 13

*Qipeng Yuan

14

[email protected]

15

Telephone: +86-10-64437610; fax: +86-10-64437610;

16

Address: 15 Beisanhuan East Road, Chaoyang District, Beijing 100029, China

1

ACS Paragon Plus Environment

ACS Synthetic Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

17

Abstract

18

Hydroxytyrosol (HT) is a valuable natural phenolic compound with strong antioxidant

19

activity and various physiological and pharmaceutical functions. In this study, we established

20

an artificial pathway for HT biosynthesis. First, efficient enzymes were selected to construct

21

tyrosol biosynthetic pathway. Aro10 from Saccharomyces cerevisiae was shown to be a

22

better ketoacid decarboxylase than Kivd from Lactococcus lactis for tyrosol production.

23

While knockout of feaB significantly decreased accumulation of the byproduct

24

4-hydroxyphenylacetic acid, overexpression of alcohol dehydrogenase ADH6 further

25

improved tyrosol production. The titers of tyrosol reached 1469 ± 56 mg/L from tyrosine and

26

620 ± 23 mg/L from simple carbon sources, respectively. The pathway was further extended

27

for HT production by overexpressing Escherichia coli native hydroxylase HpaBC. To

28

enhance transamination of tyrosine to 4-hydroxyphenylpyruvate, NH4Cl was removed from

29

the culture media. To decrease oxidation of HT, ascorbic acid was added to the cell culture.

30

To reduce the toxicity of HT, 1-dodecanol was selected as the extractant for in situ removal of

31

HT. These efforts led to an additive increase in HT titer to 1243 ± 165 mg/L in the feeding

32

experiment. Assembly of the full pathway resulted in 647±35 mg/L of HT from simple

33

carbon sources. This work provides a promising alternative for sustainable production of HT,

34

which shows scale up potential.

35

KEYWORDS: hydroxytyrosol; microbial synthesis; shikimate pathway; biphasic cultivation;

36

4-hydroxyphenylacetic acid 3-hydroxylase

2

ACS Paragon Plus Environment

Page 2 of 29

Page 3 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Synthetic Biology

37

Phenolic compounds constitute the biggest group of natural antioxidants. They have drawn

38

much attention due to their diverse biological activities and beneficial health effects.

39

Hydroxytyrosol (3, 4-dihydroxyphenylethanol or HT) is a natural phenolic compound present

40

mainly in olives. It exhibits powerful antioxidant activity and contributes to the beneficial

41

properties of virgin olive oil.1 Numerous studies have demonstrated the disease-preventing

42

potential of HT. It can protect blood lipids from oxidative damage,2 inhibit platelet

43

aggregation,3

44

anti-inflammatory6 and anti-microbial activities.7 In addition, HT has good bioavailability

45

and no known toxic effects.8,

46

supplements, functional foods and even medicine. However, due to the lack of efficient

47

production methods, this valuable compound is not commercially available in large scale.

48

HT naturally exists in fruits and vegetables, but it is mostly abundant in olive trees (Olea

49

europaea) as a product of oleuropein degradation. Consequently, olive tree derivatives are the

50

main accessible sources for HT. Processes have been developed to extract HT from fresh

51

olive leaves 10, olive pomace 11 or olive mill wastewaters (OMWWs).12 Although the starting

52

materials are cheap and abundant, these processes have several drawbacks, such as use of

53

strongly-acid aqueous steam, low recovery yields and long duration.13 Chemical synthesis

54

processes have been developed using structure analogues of HT, such as tyrosol and

55

2,3-dihydroxybenzaldehyde as the starting materials.14,

56

relatively expensive and the processes often require protection and de-protection steps,

57

lowering the overall yields.13 Besides chemical conversion, bioconversion studies have also

58

been conducted to produce HT from precursors like tyrosol, 2-phenylethanol and

and

scavenge

free

9

radicals.4

It

also

exhibits

anti-carcinogenic,5

Therefore, it possesses promising applications in food

15

3

ACS Paragon Plus Environment

However, these substrates are

ACS Synthetic Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

59

3-nitrophenethyl alcohol (3NPA). In one study, mushroom tyrosinase was used to catalyze

60

the hydroxylation of tyrosol to HT while ascorbic acid was used as a reducing agent to

61

prevent over-oxidation of HT.16 However, tyrosinase is instable and expensive and its activity

62

is inhibited by both the phenols and ascorbic acid. Some aromatic compound degrading

63

microorganisms, such as Serratia marcescens, Pseudomas aeruginosa, Pseudomonas putida

64

F6, Halomonas sp. strain HTB24 were used to convert tyrosol to HT by their native

65

hydroxylases.17-20 4-Hydroxyphenylacetic acid 3-hydroxylase was demonstrated to be

66

responsible for tyrosol hydroxylation.21 In another study, the activity of toluene

67

4-monooxygenase (T4MO) was improved by a combination of directed evolution and

68

rational design for HT production from 2-phenylethanol by two successive oxidation

69

reactions.22 An engineered nitrobenzene dioxygenase (NBDO) was also used for HT

70

production from 3NPA.23 Both T4MO and NBDO suffer from poor regio-specificity and low

71

enzyme activity.

72

A major obstacle to bioconversion production of HT is the cost of the substrates. The advent

73

of metabolic engineering and synthetic biology provides new opportunities and great

74

potential to tackle this obstacle by reconstituting natural pathways or even designing artificial

75

pathways.24 In fact, efficient production of a variety of aromatic compounds, such as

76

5-hydroxytryptophan25, caffeic acid26

77

carbon sources using metabolic engineered microorganisms. To date, only one pathway has

78

been reported for microbial production of HT.28 This pathway starts from E. coli native

79

metabolite tyrosine. Tyrosine is converted to HT by the sequential catalysis of tyrosine

80

hydroxylase (TH), L-Dopa decarboxylase, tyramine oxidase and native dehydrogenases. The

and flavoniods27 has been achieved from simple

4

ACS Paragon Plus Environment

Page 4 of 29

Page 5 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Synthetic Biology

81

first two enzymes are from mammalians and their encoding genes were codon-optimized for

82

better expression in E. coli. In addition, a cofactor regeneration pathway was constructed to

83

support the activity of TH. With strain and pathway optimization, the final strain produced

84

0.08 mM (12.3 mg/L) HT from glucose.28 This research opens the possibility for HT

85

production from simple carbon sources. However, the pathway efficiency need to be further

86

improved for economical production of HT.

87

Considering the wide potential applications of HT and the lack of economical feasible

88

production methods, in this study we designed a novel artificial pathway. HT was produced

89

from tyrosine or from simple carbon sources using metabolic engineered E. coli and under

90

optimized conditions the titers reached 1243 ± 165 mg/L and 647±35 mg/L, respectively.

91

Results and Discussion

92

Design an artificial pathway for HT biosynthesis

93

HT is a natural product of oleuropein degradation. Previously, HT has been produced from

94

tyrosol via whole cell bioconversion.21 The conversion of tyrosol to HT requires only one

95

hydroxylation step and 4-HPA 3-hydroxylase (HpaBC) from P. aeruginosa was demonstrated

96

to be able to carry out this reaction. HpaBC is widely distributed in many microorganisms

97

including E. coli. E. coli HpaBC has broad substrate activity and has been used for

98

hydroxylation of a variety of aromatic compounds, including coumaric acid,26 monolignols29

99

and flavonoids30. In this study, E. coli HpaBC was selected to perform the hydroxylation of

100

tyrosol to HT. We purified the enzyme and characterized its catalytic parameters toward

101

tyrosol (km= 18.1 ± 0.7 µM, Kcat = 12.1 ± 0.1 min-1) following the protocol described by Lin et.al30.

102

As tyrosol is a relative expensive precursor, it is necessary to extend upward to achieve HT

5

ACS Paragon Plus Environment

ACS Synthetic Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

103

production from a cheaper substrate or ultimately from simple carbon sources such as glucose

104

and glycerol. To this end, an efficient tyrosol biosynthesis pathway is required to connect

105

with the downstream conversion step. So far, two pathways have been reported for tyrosol

106

biosynthesis. In the first pathway, tyrosine is converted to tyrosol by the consecutive catalysis

107

of tyrosine decarboxylase, tyramine oxidase and alcohol dehydrogenase (ADH).31 In the

108

second pathway, tyrosol, as an intermediate of salidroside biosynthesis, was produced from

109

4-hydroxyphenylpyruvate (4-HPP) through the sequential catalysis of ketoacid decarboxylase

110

(KDC) and ADH.32 Judging from the obtained titers of tyrosol, the efficiency of the latter

111

pathway is significantly higher than that of the former one. Accordingly, an artificial pathway

112

was designed for HT biosynthesis (Figure 1).

113

Tyrosol production from tyrosine

114

Feeding experiment was carried out to determine the optimal enzyme combination for tyrosol

115

production. 4-HPP is an intermediate in tyrosine biosynthesis and E. coli native

116

aminotransferases such as TyrB catalyze the reversible transformation between tyrosine and

117

4-HPP. Considering the easier commercial availability, tyrosine instead of 4-HPP was chosen

118

as the substrate for the bioconversion study. Two KDCs, Kivd from L. lactis and Aro10 from

119

S. cerevisiae, were selected as the candidates.

120

At first, E. coli BW25113 was transformed with plasmid pZE-Kivd, generating strain BK,

121

and used for the feeding experiment. In 48 h, BK produced only 198 ± 9 mg/L of tyrosol, but

122

accumulated 450 ± 35 mg/L of 4-HPA (Figure 2A&B). In E. coli, FeaB is the main

123

dehydrogenase that converts 4-HPAA to 4-HPA. To reduce 4-HPA production, gene feaB in

124

the chromosome was knocked out, generating strain BW∆feaB. BW∆feaB was transformed

6

ACS Paragon Plus Environment

Page 6 of 29

Page 7 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Synthetic Biology

125

with plasmid pZE-Kivd, generating strain B∆K. Strain B∆K produced 499 ± 8 mg/L of

126

tyrosol while the amount of 4-HPA was significantly decreased to 36 ± 2 mg/L. To test

127

whether native ADHs are sufficient to reduce 4-HPAA to tyrosol, ADH6 with broad substrate

128

ranges from S. cerevisiae was co-expressed with Kivd on plasmid pZE-Kivd-ADH6. Strain

129

BKA produced 600 ± 23 mg/L of tyrosol but still accumulated 293 ± 12 mg/L of 4-HPA. In

130

comparison, strain B∆KA produced 778 ± 33 mg/L of tyrosol with only 30 ± 3 mg/L of

131

4-HPA (Figure 2A&B). It was observed that the cell growth decreased with the increase of

132

tyrosol production (Figure 2C), which was generally coincident with the trend of the

133

inhibition experiment (Figure S1). Aro10 was shown to have better activity toward

134

phenylpyruvate than Kivd.33 Since 4-HPP is very similar with phenylpyruvate, we assumed

135

that Aro10 may be a better candidate than KivD for tyrosol production. We then replaced

136

Kivd with Aro10, generating plasmid pZE-Aro10-ADH6. In 48 h, strain B∆AA carrying this

137

plasmid produced 1469 ± 56 mg/L of tyrosol along with 117 ± 13 mg/L of 4-HPA at 37 ℃.

138

We also investigated the effect of cultivation temperature on the conversion efficiency. At

139

30 ℃, strain B∆AA produced 692 ± 20 mg/L of tyrosol, which is less than half of that at 37 ℃

140

(Figure 2D).

141

From the results obtained above, we drew several conclusions. Firstly, disrupting feaB can

142

significantly decrease 4-HPA accumulation, but there are still other minor aldehyde

143

dehydrogenases that can act on 4-HPAA. Secondly, although disrupting feaB is sufficient to

144

decrease 4-HPA production, overexpressing ADH6 is further beneficial to tyrosol production.

145

Thirdly, Aro10 performs better than Kivd for tyrosol production. Fourthly, 37 ℃ is better

146

than 30 ℃ for the bioconversion.

7

ACS Paragon Plus Environment

ACS Synthetic Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

147

Tyrosol production from simple carbon sources

148

After achieving efficient production of tyrosol from tyrosine, we further investigated de novo

149

production of tyrosol. To this end, BW∆feaB was transformed with plasmids pZE-Aro10-

150

ADH6 and pCS-TPTA, generating strain B∆AAT. Plasmid pCS-TPTA is a medium copy

151

plasmid overexpressing four key enzymes to boost the carbon flux through shikimate

152

pathway. In shake flask experiment cell growth of B∆AAT peaked at 24 h, but tyrosol titer

153

kept increasing till 48 h, reaching 550 ± 26 mg/L (Figure 3A). We also tested the capability of

154

another strain QH4∆feaB for tyrosol production. Strain QH4 is a derivative of the

155

phenylalanine overproducer ATCC31884 and has been successfully used for microbial

156

production of a variety of aromatic compounds.26, 29, 34 Transforming the same two plasmids

157

into QH4∆feaB generated strain Q∆AAT. Strain Q∆AAT produced 620 ± 23 mg/L of tyrosol

158

in 48 h (Figure 3B). As for 4-HPA accumulation, strains B∆AAT and Q∆AAT showed

159

different trends. During the cultivation process, B∆AAT accumulated less than 40 mg/L of

160

4-HPA (Figure 3A). However, the titer of 4-HPA produced by strain Q∆AAT increased to 273

161

± 6 mg/L during the first 24 hours, and then decreased to 86 ± 5 mg/L at 48 h (Figure 3B).

162

HT production from tyrosine

163

To extend tyrosol pathway for HT production, E. coli native hydroxylase HpaBC was

164

overexpressed on plasmid pCS-HpaBC. Strain BW∆feaB was transformed with

165

pZE-Aro10-ADH6 and pCS-HpaBC, generating strain B∆AAH, for bioconversion of HT

166

from tyrosine. Strain B∆AAH was able to efficiently convert tyrosine to HT. HT titer peaked

167

at 24 h, reaching 710 ± 34 mg/L. Unlike that of tyrosol, the titer of HT decreased

168

significantly with time after 24 h. At 48 h, only 261 ± 23 mg/L of HT remained in the culture

8

ACS Paragon Plus Environment

Page 8 of 29

Page 9 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Synthetic Biology

169

(Figure 4A). This indicates that introduction of an additional hydroxyl group makes HT less

170

stable than tyrosol. We also observed that the color of the cell cultures turned black. This is

171

because HpaBC can also act on tyrosine and the product L-Dopa is unstable and readily

172

oxidized to melanin. To circumvent this problem, in our previous study a co-culture system

173

was designed to prevent the accessibility of HpaBC to tyrosine.29 Here, the problem was

174

alleviated by simply removing NH4Cl in the M9 medium. The removal of NH4Cl is supposed

175

to enforce the cells to use tyrosine as an alternative nitrogen source, which can provide a

176

driving force for transamination of tyrosine to 4-HPP. As expected, using the modified M9

177

medium lacking NH4Cl HT titer peaked at 36 h and reached 831 ± 49 mg/L, although the cell

178

density was decreased by 20 %. Moreover, ascorbic acid was added to the cell culture to

179

decrease HT oxidation. This effort further improved the titer to 972 ± 160 mg/L (Figure 4A).

180

HT showed antibacterial activity against several bacterial strains.35 To test its inhibition effect

181

on E. coli BW25113, HT was added to the cell cultures immediately after inoculation to the

182

final concentrations ranging from 0-2000 mg/L. The cell growth of BW25113 was

183

significantly inhibited at concentrations higher than 1000 mg/L. The final OD600 at 2000

184

mg/L of HT was about 40 % of that of the control (Figure 5A).

185

To alleviate the toxic effect of HT, the biphasic process was used to achieve in situ removal of

186

HT from the culture media. A key step to biphasic production is to choose a suitable organic

187

solvent with good biocompatibility and product affinity. Ethyl acetate is the most commonly

188

used solvent to extract phenolic compounds including HT from OMWWs under acidic

189

conditions36. However, this solvent is highly volatile and toxic to the cells. Therefore, we

190

chose hexyl acetate instead of ethyl acetate as a candidate solvent. Besides, 1-octanol has

9

ACS Paragon Plus Environment

ACS Synthetic Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

191

been used as an extractant for phenol production by solvent-tolerant host P. putida S12,

192

leading to a twofold increase in phenol titer.37 In another study, 1-dodecanol was proved to be

193

a promising solvent for 3-methylcatechol production in a biphasic partitioning bioreactor 38.

194

As HT is also a phenolic compound, it is expected that the longer-chain alcohols could be

195

suitable solvents for HT extraction.

196

A limitation accompanied with organic solvents is their potential toxicity towards the

197

biocatalyst. To test the toxicity of the three extractants, they were added to the cell cultures

198

with the O/W ratio of 10%. It was shown that hexyl acetate and 1-octanol exhibited severe

199

inhibition on cell growth while 1-dodecanol showed only minor inhibition (Figure 5B). We

200

also measured the partition coefficients of HT between the extractants and the M9 medium.

201

As indicated in Figure 5C, these organic solvents showed modest capability to extract HT

202

from the culture medium. Although 1-octanol has the highest partition coefficient,

203

1-dodecanol was used for further biphasic extraction study due to its good biocompatibility.

204

For this, 25 mL of 1-dodecanol was added to 50 mL of the cell cultures at 12 h after

205

inoculation. Adding 1-dodecanol, together with removal of NH4Cl led to the production of

206

1037 ± 108 mg/L of HT at 36 h. By the combination of adding1-dodecanol and ascorbic acid,

207

and removing NH4Cl, the final titer of HT reached 1243 ± 165 mg/L, which is 75 % higher

208

than that in the original medium (Figure 4A). Besides, it is observed that the biphasic

209

cultivation also led to increased accumulation of the intermediate tyrosol (Figure 4B).

210

HT production from simple carbon sources

211

To realize de novo production of HT, E. coli BW∆feaB and QH4∆feaB were transformed

212

with plasmids pZE-Aro10-ADH6 and pCS-TPTA-HpaBC, generating strain B∆AATH and

10

ACS Paragon Plus Environment

Page 10 of 29

Page 11 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Synthetic Biology

213

Q∆AATH. Strain B∆AATH produced 507±35 mg/L of HT at 32 h. Adding ascorbic acid

214

improved the titer to 596±63 mg/L at 48 h. Adding both ascorbic acid and 1-dodecanol led

215

to the production of 647±35 mg/L of HT at 48 h (Figure 6). Tyrosol and 4-HPA were

216

accumulated only in tiny amount. However, when strain Q∆AATH was used for HT

217

production, the titer reached only 395±25 mg/L at 48 h (Figure 7A). Meanwhile, both

218

tyrosol and 4-HPA were observed to be accumulated in significant amount (Figure 7B&C).

219

Like the previous observation, the titer of 4-HPA peaked at 24 h and decreased thereafter.

220

Therefore, BW∆feaB is more suitable than QH4∆feaB for HT production.

221

Discussion

222

In this work, we designed a novel biosynthetic pathway and achieved efficient HT production

223

from simple carbon sources. HT titer reached 647±35 mg/L, which is over 50 times higher

224

than that obtained previously.28 The product titer largely relies on the performance of pathway

225

enzymes. While several mammalian enzymes were involved in the previous pathway, the

226

enzymes used in this work are all from microorganisms. Generally, microbial enzymes are

227

more compatible with microbial hosts, because many mammalian enzymes require

228

posttranslational modifications and inner membrane structures for their activity which is

229

lacking in microorganisms.

230

Tyrosol is a key intermediate in HT biosynthesis. Microbial production of tyrosol had been

231

reported before.32,

232

production and made several new observations. We noticed that disruption of feaB can

233

significantly decrease but cannot completely eliminate 4-HPA accumulation especially under

234

the genetic background of QH4, indicating there are other aldehyde dehydrogenases that can

39

As a key step toward HT production, we also optimized tyrosol

11

ACS Paragon Plus Environment

ACS Synthetic Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 29

235

oxidize 4-HPAA. Besides, in previous studies 4-HPAA was reduced to tyrosol by E. coli

236

native dehydrogenases. In this work, we found that overexpression of a dehydrogenase is

237

beneficial to tyrosol production. Moreover, exogenous aromatic amino acid transaminase had

238

been overexpressed in E. coli to convert tyrosine to 4-HPP.39 We observed that E. coli native

239

transaminase is sufficient to carry out the transamination. Eliminating NH4Cl could provide a

240

driving force and enhance the transamination.

241

As an antioxidant, HT is easy to be oxidized during the production process. Adding ascorbic

242

acid to the culture media was shown to be beneficial for HT production. HT shows inhibition

243

effect on cell growth. To circumvent this problem, biphasic cultivation was adopted for in situ

244

removal of HT. Among the three extractants selected, 1-dodecanol was shown to be the

245

optimal one with good biocompatibility and modest extraction capability. Ethyl acetate had

246

been used as the solvent to extract HT from OMWWs under acidic conditions

247

Acidification can decrease the dissociation of HT in the water system and improve the

248

partition coefficient. However, the cell growth will be retarded under lower pH, which

249

generates a dilemma between high partition coefficient and good cell growth. Even with

250

modest partition coefficient, the biphasic system was able to significantly improve HT

251

production. In further work, other strategies may be adopted to further alleviate the toxicity

252

and facilitate the recovery of HT, e.g., introducing acyl transferase to further convert HT into

253

its esters.

254

Materials and methods

255

Bacterial strains and culture media.

256

E. coli XL1-Blue (from Stratagene) was used for plasmid construction and propagation. E.

12

ACS Paragon Plus Environment

36

.

Page 13 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Synthetic Biology

257

coli BW25113 (from Coli Genetic Stock Center) and QH4 were used for production of

258

tyrosol and HT. LB medium was used for seed culture. Modified M9 medium was used for

259

feeding experiments and de novo production of tyrosol and HT. LB medium contains 10 g/L

260

tryptone, 5 g/L yeast extract, and 10 g/L NaCl. The modified M9 medium contains 2 g/L

261

MOPS (morpholinepropanesulfonic acid), 10 g/L glycerol, 2.5 g/L glucose, 6 g/L Na2HPO4,

262

0.5 g/L NaCl, 3 g/L KH2PO4, 2 g/L NH4Cl, 1mM MgSO4, 0.1 mM CaCl2 and yeast extract.

263

When needed, ampicillin, kanamycin and ascorbic acid were added to the medium at 100

264

µg/mL, 50 µg/mL, and 1 g/L, respectively.

265

DNA manipulation. Plasmids pZE12-luc (high copy), pCS27 (medium copy) were used for

266

pathway assembly. Plasmids pCS-HpaBC, pCS-TPTA were constructed as described in our

267

previous study.29,

268

genome DNA. Gene encoding Kivd was amplified from L. lactis genome DNA. Gene Kivd

269

was cloned into pZE12-luc using KpnI and XbaI, generating pZE-Kivd. Plasmid

270

pZE-Kivd-ADH6 was constructed by inserting Kivd and ADH6 encoding genes into

271

pZE12-luc using KpnI, PstI and XbaI. Plasmid pZE-Aro10 was constructed by inserting

272

Aro10 encoding gene into pZE12-luc using KpnI and SphI. Plasmid pZE-Aro10-ADH6 was

273

constructed by inserting Aro10 and ADH6 encoding genes into pZE12-luc using KpnI,

274

BamHI and XbaI. Plasmid pCS-TPTA-HpaBC was constructed by inserting the expressing

275

cassette PLlacO1-HpaBC into SpeI/SacI sites of pCS-TPTA. Gene feaB on the chromosome

276

of BW25113 and QH4 was inactivated following the standard protocol of RED

277

recombination as described previously41 . Plasmids and strains used in this study are listed in

278

Table1. Primers used in this study were listed in Table S1.

40

Genes encoding Aro10 and ADH6 were amplified from S.cerevisiae

13

ACS Paragon Plus Environment

ACS Synthetic Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

279

Feeding experiments. Feeding experiments were conducted to examine the production of

280

tyrosol and HT from tyrosine. Single colonies were inoculated into 4 mL of LB media

281

containing appropriate antibiotic(s) and cultured overnight at 37 ℃. Overnight cultures (1

282

mL) were inoculated into 50 mL of M9Y media containing 5 g/L yeast extract and

283

appropriate antibiotic(s). Cells were grown at 37 ℃ and induced with 0.5 mM

284

isopropyl-β-D-thiogalactoside (IPTG). The induced strains were fed with tyrosine at 0 h, 12 h

285

and 24 h after inoculation (each time at 1 g/L). Samples were taken at several time points.

286

Optical densities at 600 nm (OD600) were measured for cell growth and the concentrations of

287

product and by-products were analyzed by HPLC. HT was extracted with ethyl acetate and

288

analyzed by ESI-MS and the molecular weight was accordance with that of the HT standard

289

(Figure S2).

290

De novo production of tyrosol and HT

291

For de novo production of tyrosol and HT, 1 mL overnight cultures were inoculated into 50

292

mL of M9Y media containing 2 g/L yeast extract and appropriate antibiotics. Cells were

293

cultivated at 37 °C and induced with 0.5 mM IPTG for 48 h. Samples were taken every 12 h.

294

The cell densities (OD600) were measured and the concentrations of the products and the

295

intermediates were analyzed by HPLC. For biphasic production of HT, 25 mL of 1-dodecanol

296

was added to 50 mL of the cell culture at 12 h after inoculation. Both the water phase and the

297

organic phase were subjected to HPLC analysis. The total titers were calculated by adding the

298

concentrations in the water phase with ½ of the concentrations in the organic phase.

299

Toxicity test and estimation of the partition coefficient

300

To test the toxicity of HT, HT was fed to the cell cultures of E. coli BW25113 to the final

14

ACS Paragon Plus Environment

Page 14 of 29

Page 15 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Synthetic Biology

301

concentrations of 0, 250, 500, 1000 and 2000 mg/L, respectively. Similarly, to test the

302

toxicity of the organic solvents, 5 mL of the individual organic solvent was added to 50 mL of

303

the cell cultures. The toxicity was evaluated by monitoring cell growth.

304

To estimate the partition coefficient of HT, M9 medium containing 500 mg/L of HT was

305

mixed thoroughly with equal volume of the organic solvent. After ten minutes’ standing,

306

samples were taken and HPLC analysis was carried out. The partition coefficient was

307

expressed as the ratio of HT concentrations in the organic and the water phases.

308

HPLC analysis. Tyrosol, 4-hydroxyphenylacetic acid (4-HPA) and HT were all purchased

309

from Aladdin Chemical Industry and used as the standards. Both the standards and samples

310

were analyzed and quantified by HPLC (HITACHI) equipped with a reverse-phase Diamonsil

311

C18 column (Diamonsil 5 µm, 250 × 4.6 mm) and UV–VIS detector. Solvent A was methanol

312

and solvent B was water with 0.1% trifluoroacetic acid. The column temperature was set to

313

28℃. The following gradient was used at a flow rate of 1 mL/min: 90 % to 40 % solvent B

314

for 18 min, 40 % to 90 % solvent B for 1 min, and 90 % solvent B for an additional 6 min.

315

Quantification were based on the peak areas at specific wavelengths (276 nm for tyrosol, 240

316

nm for 4-HPA and 280 nm for HT).

317 318

Author information

319

Corresponding authors

320

Email: [email protected]. Phone: +86-10-64431557

321

Email:[email protected]. Phone: +86-10-64437610

322

15

ACS Paragon Plus Environment

ACS Synthetic Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

323

Author Contributions

324

XL, XS and QY conceived the study and wrote the manuscript. XL, ZC, and YW performed

325

the experiments. XS, QY and YY directed the research and revised the manuscript.

326 327

Notes: A patent application about this technology has been filed by Beijing University of

328

Chemical Technology.

329 330

Acknowledgments

331

The authors would like to acknowledge financial support of the National Natural Science

332

Foundation of China (21606012, 21636001 and 21776008) and the Fundamental Research

333

Funds for the Central Universities (buctrc201613).

334 335

Supporting Information. Table S1 showing primers used in this study; Figure S1 showing

336

cell growth of E. coli BW25113 at different concentrations of tyrosol; Figure S2 showing

337

ESI-MS results of hydroxytyrosol.

338 339

16

ACS Paragon Plus Environment

Page 16 of 29

Page 17 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

340 341 342 343 344 345 346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383

ACS Synthetic Biology

References 1.

Visioli, F., Bellomo, G., and Galli, C. (1998) Free radical-scavenging properties of olive oil polyphenols, Biochem. Biophys. Res. Commun. 247, 60-64.

2.

Raederstorff, D. (2009) Antioxidant activity of olive polyphenols in humans: a review, Int. J. Vitam. Nutr.

3.

Petroni, A., Blasevich, M., Salami, M., Papini, N., Montedoro, G. F., and Galli, C. (1995) Inhibition of

Res. 79, 152-165. platelet aggregation and eicosanoid production by phenolic components of olive oil, Thromb. Res. 78, 151-160. 4.

Obied, H. K., Karuso, P., Prenzler, P. D., and Robards, K. (2007) Novel secoiridoids with antioxidant activity from Australian olive mill waste, J. Agric. Food Chem. 55, 2848-2853.

5.

Anter, J., Tasset, I., Demyda-Peyras, S., Ranchal, I., Moreno-Millan, M., Romero-Jimenez, M., Muntane, J., Luque de Castro, M. D., Munoz-Serrano, A., and Alonso-Moraga, A. (2014) Evaluation of potential antigenotoxic, cytotoxic and proapoptotic effects of the olive oil by-product "alperujo", hydroxytyrosol, tyrosol and verbascoside, Mutat. Res., Genet. Toxicol. Environ. Mutagen. 772, 25-33.

6.

Takeda, Y., Bui, V. N., Iwasaki, K., Kobayashi, T., Ogawa, H., and Imai, K. (2014) Influence of olive-derived hydroxytyrosol on the toll-like receptor 4-dependent inflammatory response of mouse peritoneal macrophages, Biochem. Biophys. Res. Commun. 446, 1225-1230.

7.

Bisignano, G., Tomaino, A., Lo Cascio, R., Crisafi, G., Uccella, N., and Saija, A. (1999) On the in-vitro antimicrobial activity of oleuropein and hydroxytyrosol, J. Pharm. Pharmacol. 51, 971-974.

8.

Manna, C., Galletti, P., Maisto, G., Cucciolla, V., D'Angelo, S., and Zappia, V. (2000) Transport mechanism and metabolism of olive oil hydroxytyrosol in Caco-2 cells, FEBS Lett. 470, 341-344.

9.

Aunon-Calles, D., Canut, L., and Visioli, F. (2013) Toxicological evaluation of pure hydroxytyrosol, Food

10.

Bouaziz, M., and Sayadi, S. (2005) Isolation and evaluation of antioxidants from leaves of a Tunisian

Chem. Toxicol. 55, 498-504. cultivar olive tree, Eur. J. Lipid Sci. Technol. 107, 497-504. 11.

Rigane, G., Bouaziz, M., Baccar, N., Abidi, S., Sayadi, S., and Ben Salem, R. (2012) Recovery of hydroxytyrosol rich extract from two-phase Chemlali olive pomace by chemical treatment, J. Food Sci. 77, C1077-1083.

12.

Kalogerakis, N., Politi, M., Foteinis, S., Chatzisymeon, E., and Mantzavinos, D. (2013) Recovery of antioxidants from olive mill wastewaters: a viable solution that promotes their overall sustainable management, J. Environ. Manag. 128, 749-758.

13.

Achmon, Y., and Fishman, A. (2015) The antioxidant hydroxytyrosol: Biotechnological production challenges and opportunities, Appl. Microbiol. Biotechnol. 99, 1119-1130.

14.

Bovicelli, P., Antonioletti, R., Mancini, S., Causio, S., Borioni, G., Ammendola, S., and Barontini, M.

15.

Zhang, Z.-L., Chen, J., Xu, Q., Rao, C., and Qiao, C. (2012) Efficient synthesis of hydroxytyrosol from 3,

(2007) Expedient synthesis of hydroxytyrosol and its esters, Synth. Commun. 37, 4245-4252. 4-dihydroxybenzaldehyde, Synth. Commun. 42, 794-798. 16.

Espín, J. C., Soler-Rivas, C., Cantos, E., Tomás-Barberán, F. A., and Wichers, H. J. (2001) Synthesis of the antioxidant hydroxytyrosol using tyrosinase as biocatalyst, J. Agric. Food Chem. 49, 1187-1193.

17.

Allouche, N., Damak, M., Ellouz, R., and Sayadi, S. (2004) Use of whole cells of Pseudomonas aeruginosa for synthesis of the antioxidant hydroxytyrosol via conversion of tyrosol, Appl. Environ. Microbiol. 70, 2105-2109.

18.

Allouche, N., and Sayadi, S. (2005) Synthesis of hydroxytyrosol, 2-hydroxyphenylacetic acid, and 3-hydroxyphenylacetic acid by differential conversion of tyrosol isomers using Serratia marcescens

17

ACS Paragon Plus Environment

ACS Synthetic Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427

strain, J. Agric. Food Chem. 53, 6525-6530. 19.

Brooks, S. J., Doyle, E. M., and O’Connor, K. E. (2006) Tyrosol to hydroxytyrosol biotransformation by immobilised cell extracts of Pseudomonas putida F6, Enzyme Microb. Technol. 39, 191-196.

20.

Liebgott, P.-P., Labat, M., Casalot, L., Amouric, A., and Lorquin, J. (2007) Bioconversion of tyrosol into hydroxytyrosol and 3, 4-dihydroxyphenylacetic acid under hypersaline conditions by the new Halomonas sp. strain HTB24, FEMS Microbiol. Lett. 276, 26-33.

21.

Liebgott, P.-P., Amouric, A., Comte, A., Tholozan, J.-L., and Lorquin, J. (2009) Hydroxytyrosol from tyrosol using hydroxyphenylacetic acid-induced bacterial cultures and evidence of the role of 4-HPA 3-hydroxylase, Res. Microbiol. 160, 757-766.

22.

Brouk, M., and Fishman, A. (2009) Protein engineering of toluene monooxygenases for synthesis of

23.

Bernath-Levin, K., Shainsky, J., Sigawi, L., and Fishman, A. (2014) Directed evolution of nitrobenzene

hydroxytyrosol, Food Chem. 116, 114-121. dioxygenase for the synthesis of the antioxidant hydroxytyrosol, Appl. Microbiol. Biotechnol. 98, 4975-4985. 24.

Sun, X., Shen, X., Jain, R., Lin, Y., Wang, J., Sun, J., Wang, J., Yan, Y., and Yuan, Q. (2015) Synthesis of chemicals by metabolic engineering of microbes, Chem. Soc. Rev. 44, 3760-3785.

25.

Lin, Y., Sun, X., Yuan, Q., and Yan, Y. (2014) Engineering bacterial phenylalanine 4-hydroxylase for microbial synthesis of human neurotransmitter precursor 5-hydroxytryptophan, ACS Synth. Bio. 3, 497-505.

26.

Huang, Q., Lin, Y., and Yan, Y. (2013) Caffeic acid production enhancement by engineering a phenylalanine over-producing Escherichia coli strain, Biotechnol. Bioeng. 110, 3188-3196.

27.

Pandey, R. P., Parajuli, P., Koffas, M. A., and Sohng, J. K. (2016) Microbial production of natural and non-natural flavonoids: pathway engineering, directed evolution and systems/synthetic biology, Biotechnol. Adv. 34, 634-662.

28.

Satoh, Y., Tajima, K., Munekata, M., Keasling, J. D., and Lee, T. S. (2012) Engineering of L-tyrosine

29.

Chen, Z., Sun, X., Li, Y., Yan, Y., and Yuan, Q. (2017) Metabolic engineering of Escherichia coli for

oxidation in Escherichia coli and microbial production of hydroxytyrosol, Metab. Eng. 14, 603-610. microbial synthesis of monolignols, Metab. Eng. 39, 102-109. 30.

Lin, Y., and Yan, Y. (2014) Biotechnological production of plant-specific hydroxylated phenylpropanoids, Biotechnol. Bioeng. 111, 1895-1899.

31.

Satoh, Y., Tajima, K., Munekata, M., Keasling, J. D., and Lee, T. S. (2012) Engineering of a tyrosol-producing pathway, utilizing simple sugar and the central metabolic tyrosine, in Escherichia coli, J. Agric. Food Chem. 60, 979-984.

32.

Bai, Y., Bi, H., Zhuang, Y., Liu, C., Cai, T., Liu, X., Zhang, X., Liu, T., and Ma, Y. (2014) Production of

33.

Atsumi, S., Hanai, T., and Liao, J. C. (2008) Non-fermentative pathways for synthesis of branched-chain

salidroside in metabolically engineered Escherichia coli, Scientific reports 4. higher alcohols as biofuels, Nature 451, 86-89. 34.

Lin, Y., Sun, X., Yuan, Q., and Yan, Y. (2014) Extending shikimate pathway for the production of muconic acid and its precursor salicylic acid in Escherichia coli, Metab. Eng. 23, 62-69.

35.

Bisignano, G., Tomaino, A., Cascio, R. L., Crisafi, G., Uccella, N., and Saija, A. (1999) On the in vitro

36.

Allouche, N., Fki, I., and Sayadi, S. (2004) Toward a high yield recovery of antioxidants and purified

antimicrobial activity of oleuropein and hydroxytyrosol, J. Pharm. Pharmacol. 51, 971-974. hydroxytyrosol from olive mill wastewaters, J. Agric. Food Chem. 52, 267-273. 37.

Wierckx, N. J., Ballerstedt, H., de Bont, J. A., and Wery, J. (2005) Engineering of solvent-tolerant

18

ACS Paragon Plus Environment

Page 18 of 29

Page 19 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

428 429 430 431 432 433 434 435 436 437 438 439

ACS Synthetic Biology

Pseudomonas putida S12 for bioproduction of phenol from glucose, Appl. Environ. Microbiol. 71, 8221-8227. 38.

Prpich, G. P., and Daugulis, A. J. (2007) Solvent selection for enhanced bioproduction of 3-methylcatechol in a two-phase partitioning bioreactor, Biotechnol. Bioeng. 97, 536-543.

39.

Xue, Y., Chen, X., Yang, C., Chang, J., Shen, W., and Fan, Y. (2017) Engineering Eschericha coli for enhanced tyrosol production, J. Agric. Food Chem. 65 (23), 4708–4714.

40.

Lin, Y., Yan,Y.. (2012) Biosynthesis of caffeic acid in Escherichia coli using its endogenous hydroxylase

41.

Datsenko, K. A., and Wanner, B. L. (2000) One-step inactivation of chromosomal genes in Escherichia

complex, Microb. Cell Fact. 11, 42. coli K-12 using PCR products, Proc. Natl. Acad. Sci. U. S. A. 97, 6640-6645.

19

ACS Paragon Plus Environment

ACS Synthetic Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

440

Page 20 of 29

Table 1. Plasmids and strains used in this study Plasmids

Feature

Source

pZE12-luc

PLlacO1, ColE ori, luc, Ampr

Ref. 36

r

pCS27

PLlacO1, P15A ori, Kan

Ref. 36

pZE-Kivd

pZE12-luc, Kivd from L. lactis

This study

pZE-Kivd-ADH6

pZE12-luc, Kivd from L. lactis and ADH6 from S. cerevisiae

This study

pZE-Aro10

pZE12-luc, ARO10 from S. cerevisiae

This study

pZE-Aro10-ADH6

pZE12-luc, ARO10 and ADH6

This study

pCS-TPTA

pCS27, tyrA, ppsA, tktA, aroGfbr from E. coli

Ref. 26

pCS-HpaBC

pCS27, hpaBC from E. coli

Ref. 26

pCS-TPTA-HpaBC

pCS27, PLlacO1-TPTA and PLlacO1-HpaBC

This study

Strains

Genotype

Source

BW25113

rrnBT14 ∆lacZWJ16 hsdR514 ∆araBADAH33 ∆rhaBADLD78

Coli Genetic Stock Center

QH4

E. coli ATCC 31884 with pheA and tyrA disrupted

Ref. 23

BW∆feaB

BW25113 with feaB disrupted

This study

QH4∆feaB

QH4 with feaB disrupted

This study

BK

BW25113 carrying plasmid pZE-Kivd

This study

B∆K

BW∆feaB carrying plasmid pZE-Kivd

This study

BKA

BW25113 carrying plasmid pZE-Kivd-ADH6

This study

B∆KA

BW∆feaB carrying plasmid pZE-Kivd-ADH6

This study

B∆AA

BW∆feaB carrying plasmid pZE-Aro10-ADH6

This study

B∆AAT

BW∆feaB carrying plasmids pZE-Aro10-ADH6 and pCS-TPTA

This study

Q∆AAT

QH4∆feaB carrying plasmid pZE-Aro10-ADH6 and pCS-TPTA

This study

B∆AAH

BW∆feaB carrying plasmid pZE-Aro10-ADH6 and pCS-HpaBC

This study

B∆AATH

BW∆feaB carrying plasmid pZE-Aro10-ADH6 and pCS-TPTA

This study

Q∆AATH

QH4∆feaB carrying plasmid pZE-Aro10-ADH6 and pCS-TPTA

-HpaBC

-HpaBC

441 442

20

ACS Paragon Plus Environment

This study

Page 21 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Synthetic Biology

443

Figure captions

444 445 446 447 448 449 450 451

Figure 1. The novel biosynthetic pathway of hydroxytyrosol. Compounds: 4-HPP, 4-hydroxyphenylpyruvate; 4-HPAA, 4-hydroxyphenylacetaldehyde; 4-HPA, 4-hydroxyphenylacetic acid. Enzymes: KDC, ketoacid decarboxylase; ADH, alcohol dehydrogenase; HpaBC, 4-hydroxyphenylacetic acid 3-hydroxylase; TyrB, aromatic-amino-acid aminotransferase; FeaB, phenylacetaldehyde dehydrogenase. Dashed arrows indicate that the branches are blocked.

452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471

Figure 2. Tyrosol production from tyrosine using different strains. Five strains were used and their genetic details are shown in Table 1. (A) Cell growth. (B) Tyrosol production. (C) 4-HPA accumulation. (D) Tyrosol production using strain B∆AA at 37 ℃ and 30 ℃. Error bars represent standard deviations of three replicates. Figure 3. De novo production of tyrosol by (A) strain B∆AAT and (B) strain Q∆AAT. Refer to Table 1 for the genetic details. Error bars represent standard deviations of three replicates. Figure 4. Optimization of bioconversion conditions for HT production from tyrosine. Strain BAAH was used for the bioconversion experiment. (A) Profiles of HT accumulation at different conditions. (B) Profiles of tyrosol accumulation at different conditions. Error bars represent standard deviations of three replicates.

Figure 5. Selection of an appropriate extractant for biphasic production of HT. (A) Effect of HT on cell frowth. (B) Effect of the extractants on cell growth. (C) Partition coefficients of HT between the extractants and the M9 medium. Error bars represent standard deviations of three replicates.

Figure 6. De novo production of HT. Strain B∆AATH was used for the experiments. Error bars represent standard deviations of three replicates.

472 473 474 475 476

Figure 7. De novo production of HT. Strain Q∆AATH was used for the experiments. (A) HT production. (B) Tyrosol accumulation. (C) 4-HPA accumulation. Error bars represent standard deviations of three replicates.

21

ACS Paragon Plus Environment

ACS Synthetic Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 1. The novel biosynthetic pathway of hydroxytyrosol. Compounds: 4-HPP, 4-hydroxyphenylpyruvate; 4-HPAA, 4-hydroxyphenylacetaldehyde; 4-HPA, 4-hydroxyphenylacetic acid. Enzymes: KDC, ketoacid decarboxylase; ADH, alcohol dehydrogenase; HpaBC, 4-hydroxyphenylacetic acid 3-hydroxylase; TyrB, aromatic-amino-acid aminotransferase; FeaB, phenylacetaldehyde dehydrogenase. Dashed arrows indicate that the branches are blocked. 94x45mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 22 of 29

Page 23 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Synthetic Biology

Figure 2. Tyrosol production from tyrosine using different strains. Five strains were used and their genetic details are shown in Table 1. (A) Cell growth. (B) Tyrosol production. (C) 4-HPA accumulation. (D) Tyrosol production using strain B∆AA at 37 ℃ and 30 ℃. Error bars represent standard deviations of three replicates.

208x156mm (300 x 300 DPI)

ACS Paragon Plus Environment

ACS Synthetic Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 3. De novo production of tyrosol by (A) strain B∆AAT and (B) strain Q∆AAT. Refer to Table 1 for the genetic details. Error bars represent standard deviations of three replicates. 241x324mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 24 of 29

Page 25 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Synthetic Biology

Figure 4. Optimization of bioconversion conditions for HT production from tyrosine. Strain BAAH was used for the bioconversion experiment. (A) Profiles of HT accumulation at different conditions. (B) Profiles of tyrosol accumulation at different conditions. Error bars represent standard deviations of three replicates. 224x292mm (300 x 300 DPI)

ACS Paragon Plus Environment

ACS Synthetic Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 5. Selection of an appropriate extractant for biphasic production of HT. (A) Effect of HT on cell frowth. (B) Effect of the extractants on cell growth. (C) Partition coefficients of HT between the extractants and the M9 medium. Error bars represent standard deviations of three replicates. 898x245mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 26 of 29

Page 27 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Synthetic Biology

Figure 6. De novo production of HT. Strain B∆AATH was used for the experiments. Error bars represent standard deviations of three replicates. 209x148mm (300 x 300 DPI)

ACS Paragon Plus Environment

ACS Synthetic Biology 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 7. De novo production of HT. Strain Q∆AATH was used for the experiments. (A) HT production. (B) Tyrosol accumulation. (C) 4-HPA accumulation. Error bars represent standard deviations of three replicates. 582x153mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 28 of 29

Page 29 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Synthetic Biology

For Table of Contents Use Only 73x44mm (600 x 600 DPI)

ACS Paragon Plus Environment