Recovery of Natural α-Ionone from Fermentation Broth - American

Mar 13, 2019 - and use natural compounds in their formulations.2−4 ...... (15) Fuganti, C.; Serra, S.; Zenoni, A. Synthesis and Olfactory. Evaluatio...
0 downloads 0 Views 796KB Size
Subscriber access provided by TULANE UNIVERSITY

Chemistry and Biology of Aroma and Taste

Recovery of natural #-ionone from fermentation broth Ilya Lukin, Guido Jach, Isabell Wingartz, Peter Welters, and Gerhard Schembecker J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.8b07270 • Publication Date (Web): 13 Mar 2019 Downloaded from http://pubs.acs.org on March 20, 2019

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 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 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.

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 28

Journal of Agricultural and Food Chemistry

1

Recovery of natural α-ionone from fermentation broth

2

Ilya Lukin1, Guido Jach2, Isabell Wingartz1, Peter Welters², Gerhard Schembecker1*

3 4

1Laboratory

of Plant and Process Design, Department of Biochemical and Chemical Engineering, TU Dortmund University, Emil-Figge-Strasse 70, D-44227 Dortmund, Germany

5

2Phytowelt

6 7

*Corresponding author: E-mail address: [email protected] Tel.: +49 (0)231 755 2338; Fax: +49 (0)231 755 2341

8

Keywords

9

natural aromas, ionone, metabolic engineering, downstream process, process development

Green Technologies GmbH, Kölsumer Weg 33, D-41334 Nettetal, Germany

10

Abstract

11

Recently, the market value of aromas has constantly been rising. As the supply from natural feedstock is

12

limited, the biotechnological production receives more interest. So far, only a few attempts have been

13

made to produce α-ionone, a valued essential aroma of raspberry, biotechnologically. This study reports

14

a production process for enantiopure (R)-α-ionone from lab scale (2 L – 150 L) with typical titer of 285

15

mg/L-broth to industrial scale (up to 10,000 L) with a titer up to 400 mg/L-broth focusing on the

16

development of a downstream process with maximized yield at minimized effort. The developed recovery

17

consists of solid-liquid extraction from the biomass at φ = 0.4 g-n-hexane/g-biomass, for 90 minutes at

18

ambient temperature and of adsorption from the aqueous supernatant at Φ = 0.5 g-DiaionHP20/mg-

19

α-ionone followed by the desorption at Ψ = 30 g-n-hexane/g-DiaionHP20. Altogether, natural α-ionone

20

could be gained in substantial quantity and purity of > 95 %.

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

21

1

Introduction

22

Over the past few decades, flavors and fragrances (F&F) or aromas have become profitable fine chemicals

23

with a continually growing market value due to their increasing use in goods like foods, feeds and

24

beverages, household products, cosmetics and toiletries, perfumery and even pharmaceuticals1.

25

Moreover, there is a strongly growing consumer-driven demand of the aroma industry to provide and use

26

natural compounds in their formulations2–4. Unfortunately, the availability of such aromas in natural

27

sources is limited, and their recovery is expensive. One way to increase the availability and by that to

28

decrease the costs of natural aromas is the application of plant biotechnology tools. Streamlining the

29

metabolite flows towards the desired path can immensely increase the aroma production in the plants.

30

The more profound understanding and control of regulatory mechanisms of secondary metabolism can

31

reduce seasonal and regional variations of aroma in the natural feedstock. Introduction of additional

32

resistances against environmental changes can lead to less crop loss. Although potent, plant biotechnology

33

is still limited by long generation times, large space needs, broad byproduct spectrum, difficult product

34

recovery, and general customer skepticism. Especially the recent developments in the modern analytics

35

and genetic engineering opened a way for biotechnological aroma production as a serious long-term

36

alternative to natural extraction5. On the one hand, a lot of bacteria and fungi are known to produce

37

commercially valuable aromas naturally4. On the other hand, the “working horses” of white biotechnology

38

like E. coli or S. cerevisae can be easily modified to produce the desired aroma compound in high quality

39

and purity. Biotechnological aroma production is rapid, highly reproducible, flexibly scalable, can be

40

carried out at mild conditions, and uses renewable substrates like sugar or even agricultural residual

41

streams. It yields an enantiopure product and does not produce any toxic waste.6,7 Besides other

42

advantages, biotechnologically derived aromas can be labeled as natural8 leading to a strong market

43

advantage for F&F companies.

44

Among various F&F compounds, the class of norisoprenoids, which are derived from the enzymatic

45

cleavage of carotenoids, is in the focus of current research. In particular, α-ionone, a key flavor of ACS Paragon Plus Environment

Page 2 of 28

Page 3 of 28

Journal of Agricultural and Food Chemistry

46

raspberry9–11 and blackberry12, is an important aromatic compound with a worldwide use on a scale of 100

47

to 1000 tons per year13. Having an extremely low odor threshold of 0.4 ppb14 - 3.2 ppb15 and warm, flowery

48

scents with a note of violet16, α-ionone is valued in cosmetics and perfumery13. As the most natural aromas,

49

α-ionone is a secondary metabolite of numerous plants, and few fungi or bacteria. Naturally, it results

50

from oxidative cleavage of precursors α-carotene or ε-carotene17. Although α-carotene is almost

51

ubiquitous, ε-carotene is present in few plants only18. In natural sources, α-ionone is found in minor

52

quantities, with concentrations ranging from some ppb to few ppm, and in combination with the

53

dominating ß-ionone, which makes its extraction from natural stock tedious and expensive9,19. Combined

54

with the concerns on supply security and regional as well as seasonal stock variations, this has led to the

55

fact, that most of α-ionone on the market is produced chemically nowadays6,7. However, chemical

56

synthesis is non-sustainable, consumes much energy, cannot be labeled natural and results in racemic

57

mixtures7. The latter, in particular, is problematic as the (S)-enantiomer of α-ionone possesses undesirable

58

scent properties making the resulting product less worth. Hence, there is a strong industry demand for

59

natural, enantiomeric pure (R)-α-ionone.

60 61

So far, there have been only a few attempts on the biochemical production of enantiopure (R)-α-ionone

62

and only one supplier of naturally labeled α-ionone could be found20. The company Phytowelt developed

63

2015 the first patent explicitly pending fermentative production system for pure (R)-α-ionone, which

64

already has been brought to technical scale21. The first ever patent, disclosing the possibility of

65

biotechnological production of carotenoids and their derivatives, as e.g. α-ionone, via fermentation of

66

genetically modified Y. lipolytica or E. coli, also dates 201522. The most recent breakthrough in the

67

biotechnological production of α-ionone has been reported 2018 by Zhang et al.23. The research group

68

designed an E. coli based “plug-and-play” system for the production of α-ionone, β-ionone, psi-ionone,

69

retinal and retinol. In small-scale (250 mL) fed-batch biphasic fermentations with isopropyl myristate as

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

70

the second organic phase for an in-situ product removal a titer of 480 mg/L α-ionone has been reported.

71

Unfortunately, substantial amounts of β-ionone and psi-ionone were found in the reported broth,

72

lowering the purity of the final product. Besides, as the most used route for the chemical synthesis of

73

α-ionone is the well-known acid catalyzed cyclisation of psi- or pseudoionone24, its presence makes the

74

final natural product indistinguishable from the chemically synthesized.

75 76

Although the fermentation of aroma compounds is quite advantageous, its recovery from crude

77

biochemical broth can be challenging. As we previously described, there is a lot of information available

78

about the lab-scale analytical aroma recovery from the fermentation broth as well as about the industrial

79

aroma recovery from plant sources. For instants, solid adsorbents were used on a lab-scale for the

80

recovery, concentration, and fractionation of aromas for several decades25,26. In the form of solid-phase

81

microextraction (SPME) adsorbents are currently still the most used lab-scale technique prior to the

82

chromatographic analytics of aromas27. In the case of industrial scale application, hydro-distillation and

83

solvent extraction are widely used for the recovery of aromas from plant material28. However, there is a

84

gap in transferring the lab-scale techniques such as adsorption to an industrial aroma recovery from the

85

fermentation broth as not all traditional aroma recovery techniques are suitable for the processing of

86

highly diluted, solid containing fermentation broth. Product deterioration, high energy consumption, or

87

high costs are only some challenges to name.29 In food and beverage processing, supercritical extraction

88

and membrane technology became state-of-the-art techniques for the recovery of volatile compounds like

89

aromas30,31. Dealing with biotechnologically produced aromas the recovery task strongly depends on the

90

product location and the phase distribution determined by the product’s physical properties like

91

hydrophobicity and volatility. The systematic development of a suitable recovery technique for the

92

biochemically produced aroma compounds can help to bridge the gap from lab-scale research to economic

93

industrial scale production.

ACS Paragon Plus Environment

Page 4 of 28

Page 5 of 28

Journal of Agricultural and Food Chemistry

94 95

In this work, enantiopure (R)-α-ionone was produced by fermentation of metabolically engineered E. coli

96

using patented technology21 comprising the use of plant-based heterologous enzyme cascade for

97

ε-carotene production and subsequent α-ionone release via carotenoid-cleavage dioxygenase activity. A

98

simple one-phase fed-batch fermentation was improved beforehand to generate a sufficient amount of

99

product containing biomass. Besides, a robust downstream process for the recovery and purification of

100

α-ionone from the fermentation broth was systematically developed with the aim to maximize the product

101

yield at the minimized effort. This research reports for the first time the combination of the optimized

102

enantiopure production and yield maximized recovery leading to an overall technical scale process for

103

sustainable production of natural (R)-α-ionone in sufficient quantity and market suitable purity.

104

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

105

2

Materials and Methods

106

2.1 Strains and fermentation

107

Phytowelt´s proprietary (R)-α-ionone producing E. coli strain EcPHY-G81 was used throughout the process

108

development to gain ionone containing biomass samples. This plasmid-free strain harbors the full enzyme

109

cascade required for ionone production in its genome and does not require the use of antibiotics or

110

inducers during the fermentation procedure. The strain expresses the genes crtB, crtE and crtI from

111

E. herbicola under the control of the proprietary promoter and terminator sequences to implement

112

required enzyme activities (geranylgeranyl-pyrophosphate synthase, phytoene synthase, and phytoene

113

desaturase) for the formation of intermediate lycopene. Additional expression of a mutated plant

114

lycopene epsilon-cyclase as well as a mutated plant carotenoid cleavage dioxygenase leads to quantitative

115

conversion of lycopene to the intermediate epsilon-carotene and then to the final product (R)-α-ionone.

116 117

Fed-batch fermentations were conducted at 26 °C over 120 h at 2 L – 150 L scale in different experiments

118

using the media and protocol described by Riesenberg et al.32. Feeding rates were adapted to achieve a

119

reduced growth rate and elongated fermentation times. Finally reached biomass concentration was in the

120

range of 60 – 70 g-CDW/L.

121

122

2.2 Phase separation

123

The phase separation of the fermentation broth was done in a laboratory centrifuge (Centrifuge 5804 R,

124

Eppendorf, Germany ) at 4500 rpm (3622 rcf) at 25 °C using Falcon™ tubes.

125 ACS Paragon Plus Environment

Page 6 of 28

Page 7 of 28

Journal of Agricultural and Food Chemistry

126

2.3 Product recovery

127

All extraction and adsorption experiments were performed on an overhead shaker (PTR-60, Grant-bio, UK)

128

at 50 rpm and ambient temperature using 15 mL Falcon™ tubes. The recovery of α-ionone during the

129

fermentation was made via extraction with ethyl acetate. The solvent was added to the fermentation

130

slurry at a phase ratio of 0.2 g-solvent/g-CWW. The two-stage extraction was carried out for 30 minutes

131

per stage. Fresh solvent was used in every extraction stage. After the phase separation, the solvents of

132

both stages were pooled and analyzed for α-ionone concentration using HPLC. During the optimization of

133

the preparative α-ionone extraction different solvents, phase ratios, temperatures, extraction times and

134

a number of stages were varied. Therefore, the fermentation broth was separated into the solid and

135

aqueous phases, and the solvent was added to the wet biomass at a given phase ratio, and the extraction

136

was carried out at defined conditions. For all experiments, the extraction was performed for 90 minutes

137

at ambient temperature. For the screening of the extraction kinetics, the extraction time was varied from

138

10 minutes to 120 minutes. For the investigations of the temperature effect, the overhead shaker was

139

placed in a temperature controlled incubator. After the extraction, the samples were centrifuged at

140

4500 rpm (3622 rcf) for 10 minutes, the phases were separated, and the solvent was analyzed for α-ionone

141

concentration using HPLC.

142 143

α-ionone was recovered from the aqueous supernatant by adsorption on solid resins. Different types of

144

adsorbents were added to the cell-free aqueous supernatant at a defined ratio and mixed at ambient

145

temperature overnight. Afterward, the samples were centrifuged at 4500 rpm (3622 rcf) for 10 minutes,

146

and the adsorbent was collected. α-ionone was desorbed by adding a solvent at a defined ratio to the

147

resin. The samples were mixed overnight. After the phase separation (4500 rpm (3622 rcf) for 10 minutes)

148

the solvent phase was collected for the α-ionone analysis using HPLC.

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

149

2.4 Analytics

150

α-ionone concentration was determined by HPLC analysis (Knauer, Germany) using a

151

NUCLEODUR™ 250x4 C8ec column (Macherey-Nagel, Germany). Acetonitrile (HiPerSolv

152

CHROMANORM®, VWR International, USA) and Millipore Water were used for elution. The

153

solvents were mixed with 1 % v/v acetic acid (GPR Rectapur 99 – 100 %, VWR Chemicals, USA)

154

and stored separately. The oven temperature was 40 °C and the following gradient was applied:

155

starting at 60 % ACN isocratic for 4 min to 100 % ACN gradient for 4 min to 100 % ACN isocratic

156

for 1 min and finally to 60 % ACN equilibration step for 6 min. The approximate total retention

157

time of α-ionone was 12.2 min. The concentration was measured by a DAD detector at 245 nm

158

(DAD Knauer, Germany) according to prior calibration with analytical standard α-ionone

159

(Sigma-Aldrich/Merck, Germany).

ACS Paragon Plus Environment

Page 8 of 28

Page 9 of 28

Journal of Agricultural and Food Chemistry

160

3

Results and discussion

161

The production strain has been consecutively optimized for the biosynthesis of the desired target product.

162

The strain used in this paper has been chosen because it proved to be remarkably robust through all stages

163

of up-scaling with highly reproducible results in all fermentation experiments. The main parameters of the

164

fermentation process like media components, feeding strategy, temperature, and duration, are commonly

165

used measures to achieve the highest product yields and economically feasible processes. These issues

166

have already been addressed beforehand and lead to the fermentative process being used in this study to

167

generate representative samples for downstream process development and evaluation.

168

3.1 Fermentative α-ionone production

169

A typical result of the fermentation of the used strain PHY-G81 at 2 – 150 L scale is given in figure 1. Cells

170

show steady growth over 123 h of fermentation time, with a somewhat slower growth after about 72 h.

171

The yield of α-ionone also increased with the time, but at a steeper slope, indicating increasing specific

172

productivity of the cells over time. Highest product titer was 282 mg/L (123 h). Further elongation of the

173

fermentation is likely to give even high titers. However, reaching the maximum volume of broth possible

174

for the given fermentation vessel limits the possible fermentation times. Scale up trials were successfully

175

accomplished and titers of 400 mg/L were reached at 10.000 L scale within 120 h fermentation time using

176

this process.

177 178

Carotenoid cleavage enzymes frequently used to release ionones from carotenoid precursor are often

179

promiscuous and accept a range of different carotenoids as substrates, which leads to the formation of

180

side-products such as pseudoionone (via cleavage of lycopene). The presence of pseudoionone in high

181

concentrations causes, in fact, severe problems, as it cannot be separated from the desired α-ionone at

182

reasonable costs and labor due to its quite similar chemical properties. Thus, the production of α-ionone ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

183

to the desired purity (> 95%) can be severely hampered. In our process, pseudoionone concentrations

184

remained typically around 1-2 % and represented no obstacle for the downstream process.

185

Thus, the upstream part already works satisfactorily at production scale. With respect to the economics of

186

the whole production process, however, extraction and purification of the desired product are of equal

187

importance to gain yield maximized production/recovery of natural α-ionone at lowest costs, which is the

188

focus of the work described here.

189

3.2 Determination of product phase distribution

190

In order to develop an industrially suitable α-ionone downstream process, knowledge about the product’s

191

phase distribution is crucial. We previously reviewed the choice of potential techniques for the recovery

192

and purification of volatile aroma compounds depending on their phase partitioning and physical

193

properties29. As a semivolatile (Hpc = 4.5x10-5 m³atm/mol33) but rather hydrophobic (logKow = 3.8534)

194

compound, α-ionone can partition between all phases of a fermentation broth. The molecule can stick to

195

the hydrophobic biomass, be diluted into the aqueous supernatant up to its saturation and can evaporate

196

into the headspace of the bioreactor. The maximum product content in the fermentation broth was

197

determined by leaching extraction of the slurry at high phase ratio. By performing the extraction of the

198

separated phases at the same conditions, the products partitioning between biomass and aqueous

199

supernatant was evaluated. Because of the limited aqueous solubility of α-ionone (csat = 0.1 g/L35), 85 –

200

92 % w/w of the molecule is found in or on the biomass leading to the higher phase specific yield (figure

201

2) while the remainder is solved in the fermentation supernatant. At fermentation conditions

202

(1 vvm, T = 26 °C) α-ionone has also been qualitatively detected in the reactor’s off-gas at around 2 – 5 %.

203

The product losses through the off-gas were, as expected, rather low ranging around 2 – 5 %. Following

204

the guidelines proposed by the WHO36 based on the boiling temperature, α-Ionone, with the boiling point

205

of 256 °C37 under standard conditions, can be classified as semi volatile organic compound with a rather

ACS Paragon Plus Environment

Page 10 of 28

Page 11 of 28

Journal of Agricultural and Food Chemistry

206

slow volatilization from aqueous media. The Henry’s law volatility coefficient of α-ionone is three orders

207

of magnitude lower than e.g. of limonene (Hpc = 1.4 – 5.8x10-2 m³atm/mol33), a common highly volatile

208

perfumes top note aroma. For that reason, the process development is focused on the product recovery

209

from the biomass. In order to maximize the yield, the possibilities of the product recovery from the

210

fermentation supernatant are explored.

211

3.3 Development of α-ionone recovery process

212

With the majority of the product sticking to the biomass (85 – 92 % w/w), a solid-liquid solvent extraction

213

seems to be the most suitable recovery technique. In general, product extraction could be done either

214

way, after the phase separation from the solid phase only, or via simultaneous extraction of both phases

215

through the application of a water immiscible solvent. Because of the biomass concentration in the

216

fermentation slurry of around cx = 60 – 70 g-CDW/L and the limited aqueous solubility of α-ionone, the

217

specific product yield of the slurry 𝑌𝑠𝑙𝑢𝑟𝑟𝑦 = 0.22 ± 0.01 mg/g-slurry is much lower than that of a solid phase 𝑝

218

𝑌𝑠𝑜𝑙𝑖𝑑 = 2.28 ± 0.1 mg/g-solids. In this case, the solvent consumption and the specific recovery costs of the 𝑝

219

slurry extraction are much higher than that of the biomass extraction. Thus, a solid-liquid phase separation

220

is applied to the fermentation broth prior to the extraction step, and α-ionone is extracted from the

221

biomass only using a suitable solvent.

222 223

3.3.1

Product recovery from the solid phase

224

The success of an extraction largely depends on solvent selection. High affinity towards and high capacity

225

for the product, broad availability and low costs, low viscosity and evaporation enthalpy are the common

226

criteria for the solvent screening38. In addition, in the case of a natural product, the solvent should meet

227

the regulations. A way to lower the experimental effort screening the organic solvents for an extraction

228

task is the use of the solvent pre-selection tool MOPASOOL© based on a systematic procedure described ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

229

by Bergs et al.39,40. The tool ranks the possible performance of the solvents from a database according to

230

the calculated hydrophobicity similarities between the desired target compound and the solvent following

231

the principle “like dissolves in like”. Restrictions to specific application areas, like for green or foodstuff

232

solvents, further reduce the number of potential solvents to screen. The tool supports the selection

233

process narrowing the range of possibilities to search for the best performing solvent, however, does not

234

predict any recoveries. The results of the solvent ranking for α-ionone extraction as well as their

235

experimental performance can be seen in figure 3. It is evident that solvent performance is estimated to

236

increase from hydrophilic to hydrophobic solvents. As the MOPASOOL© tool is limited to organic solvents

237

only, some hydrophobic plant oils and alcohols were added to the screening.

238 239

In good agreement with the theoretical preselection, the most non-polar solvent n-hexane showed one of

240

the highest recoveries. The medium polar acetone, however, performed surprisingly well with the highest

241

yield on average. As n-hexane is a non-polar solvent, it does not extract hydrophilic components (all above

242

residual water) from the biomass leading to high purity of the extract and is therefore chosen for α-ionone

243

extraction from the biomass. Although the plant-derived solvents are ecologically friendly, sustainable,

244

and are considered to be food grade, their low extraction performance will not lead to an economic

245

α-ionone production process. Especially the high viscosity of plant oils may lead to diffusion limitations of

246

the product during the extraction limiting the yield. Besides, they tend to oxidize fast generating

247

unpleasant off-flavors41 and may be difficult to separate from the product.

248 249

In order to further maximize the product yield and to minimize the recovery costs, the phase ratio and the

250

number of the extraction stages were varied on a narrow scale each parameter at a time as the interactions

251

between solvent phase ratio and stage number in a cross-flow extraction is less important as e.g. in

252

countercurrent operation. Cross-flow leaching experiments indicate that the most α-ionone is extracted ACS Paragon Plus Environment

Page 12 of 28

Page 13 of 28

Journal of Agricultural and Food Chemistry

253

after two stages already at ϕ = 0.5 g-solvent/g-sample. Further reduction of the solvent phase ratio has

254

led to the minimum ϕ = 0.4 g-solvent/g-sample. Up to 82 % w/w of the product could be extracted at the

255

first stage and the accumulated yield of the first two stages was 96 – 98 % w/w (figure 4). The product’s

256

purity was independent of the solvent phase ratio and the number of stages by 98 % peak purity (figure 5).

257

The only detected byproduct was pseudoionone, the non-cyclic equivalent of α-ionone derived from the

258

cleavage of ε-carotene precursor lycopene. Pseudoionone has very similar physical and chemical

259

properties to α-ionone which leads to equal partitioning between the biomass and the extraction solvent.

260 261

For an increased space-time-yield of the process, the extraction time should be minimized as well. The

262

measurement of the extraction kinetics showed that the minimal extraction time of α-ionone from the

263

biomass should be 90 minutes (figure 6). Although α-ionone has a high affinity towards hydrophobic

264

n-hexane, the extraction kinetics seem to be rather slow. The main reason lies presumably in the diffusion

265

limitations of α-ionone from the cells to the solvent phase. It is assumed, that α-ionone molecules are

266

incorporated into the hydrophobic cell membrane which presents an additional diffusion resistance.

267

Moreover, the biomass coming from the solid-liquid separation has a high content of the residual

268

moisture. Because n-hexane is immiscible with water, α-Ionone needs to diffuse through a hydrate shell

269

of the cells to the bulk solvent phase. In addition, in the presence of the hydrophobic solvent, wet cells

270

tend to form agglomerates minimizing the mass transfer area and slowing the extraction kinetics. Vigorous

271

mixing using an impeller should help to overcome such limitations.

272 273

Usually, elevated extraction temperatures help to increase product’s diffusion towards solvent bulk phase.

274

It is often expected, that effect of temperature increase has a positive interaction with the solvent phase

275

ratio. The thermodynamic portioning coefficient of the target compound should increase at elevated

276

temperature helping to further reduce the solvent consumption. In order to evaluate the single factor ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

277

effect, the temperature was varied at three levels while all other parameters previously optimized were

278

held constant (figure 7). Surprisingly, here the temperature had no visible effect neither on the α-ionone

279

yield nor on extract purity. The extraction equilibrium seems to be strongly shifted towards extract by the

280

use of highly hydrophobic solvent n-hexane. As the solvent loss increases with increased extraction

281

temperature, it is recommended to perform the extraction at the ambient temperature.

282 283

In summary, the optimal conditions for α-ionone extraction from the biomass at which maximum yield

284

could be achieved with minimum solvent consumption, minimum temperature control and in lowest

285

process time are estimated as follows: n-hexane at ϕ = 0.4 g-solvent/g-biomass for 90 minutes at ambient

286

temperature. The product gained had a market-ready purity of > 95% (figure 5). As the applied solvent n-

287

hexane is considered food grade under both the EU and the US legislation with the maximum solvent

288

residue in the final product less than 1 – 30 ppm (EU)42 and less than 5 – 25 ppm (US)43, the obtained

289

α-ionone can be labeled natural food grade. In order to increase the overall α-ionone yield of the

290

downstream process, additional 8 – 15 % w/w of the product should be recovered from the aqueous phase

291

too.

292 293

3.3.2 Product recovery from the aqueous phase

294

For the recovery of α-ionone from the aqueous supernatant of the fermentation broth, liquid-liquid

295

extraction, adsorption on synthetic resins, distillation, pervaporation, and even strip-absorption could be

296

applied. As α-ionone is a high-boiling compound (Tb = 256 °C 44) at a low concentration in an aqueous

297

matrix, large amounts of low-boiling water will be evaporated during distillation or pervaporation making

298

the recovery process expensive. Stripping of a semivolatile product is indeed possible but will be tedious

299

and time-consuming as the volatilization will be slow. Liquid-liquid extraction is possible and a water-

300

immiscible high hydrophobic solvent as n-hexane could be applied. Taking into account that the aqueous ACS Paragon Plus Environment

Page 14 of 28

Page 15 of 28

Journal of Agricultural and Food Chemistry

301

phase takes 95 % w/w of the fermentation broth but contains maximum 15 % w/w of the product, the

302

solvent consumption and by that the specific costs of the liquid-liquid extraction will be high. The specific

303

yield of the aqueous supernatant is with 𝑌𝑎𝑞𝑢𝑒𝑜𝑢𝑠 = 0.03 ± 0.0 mg/g-phase extremely low. Adsorption of α𝑝

304

ionone on synthetic resins may help to reduce the volume of the product containing stream and by that

305

the solvent consumption and the costs of the recovery. As it is often the case, at higher concentrations

306

aromas may have an inhibitory effect on the producing microorganism. By continuously removing the

307

product from the liquid phase in-line, the equilibrium can be shifted towards production increasing the

308

overall productivity of the process. Figure 8 shows the results of the screening of selected adsorbents for

309

the batch recovery of α-ionone from the aqueous supernatant of a fermentation broth. As the adsorption

310

is driven by the available binding surface, adsorbents were added at varied mass providing equal surface

311

area. Commonly available non-polar XAD resins are expected to deliver the best performance adsorbing a

312

rather hydrophobic target. A polar Diaion HP20, as well as universal activated carbon adsorbent, were

313

added to prove the hypothesis.

314 315

The analysis of the liquid phase after desorption showed that at ψ = 5 m²/ml-broth the fermentation broth

316

had been completely depleted of α-ionone. It is entirely adsorbed by all resins provided. However, the

317

amount of the product that could be recovered depended on the adsorbents used, with the recovery been

318

limited by the desorption step. Stronger hydrophobic resins like XAD showed a lower amount of recovered

319

product. Even when hydrophobic n-hexane is used as a desorption solvent, the strong binding of α-ionone

320

to the resin cannot be overcome. The best desorption rates could be achieved with a least hydrophobic

321

resin Diaion HP20. It was necessary to immensely increase the phase ratio of n-hexane up to

322

Φ = 30 g-solvent/g-resin (figure 9) to reach full desorption.

323

Desorption might be additionally limited by the residual moisture from the adsorption step. The hydrate

324

shell around the resins might present an additional diffusion resistance for the mass transfer of α-ionone ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

325

from the resin to the bulk solvent phase. For an industrial application, product desorption might be done

326

using a gradient elution starting with a semi-polar solvents like alcohol and switching to n-hexane for full

327

desorption.

328 329

In summary, α-ionone can be adsorbed from the aqueous supernatant of the fermentation broth with

330

Diaion HP20 resin. In a batch experiment, it was necessary to provide at least ψ = 5 m²/ml-broth of binding

331

surface or ψ* = 0.5 g-resin/mg-α-ionone for the maximized yield. In order to reach a full recovery of bound

332

α-ionone from the resin at least Φ = 30 g-solvent/g-resin n-hexane was needed. Due to the high

333

hydrophobicity of n-hexane the obtained product contained only α-ionone and pseudoionone at the same

334

ratio as in the aqueous supernatant of the fermentation broth.

335

336

4

Discussion

337

Besides the usual optimization of biosynthetic capacities of the production strain and the fermentation

338

process parameters the establishment of best-suited procedures for the downstream processing of the

339

desired product (R)-α-ionone is crucial to gain a complete yield maximized production process.

340

Upregulation of lycopene cyclase leads to the faster formation of ε-carotene reducing the amount of

341

residual lycopene and with that of the unwanted byproduct pseudoionone. The developed process

342

presents a way for sustainable production of sufficient quantity of natural enantiopure (R)-α-ionone in

343

market suitable purity. The market price of chemical α-ionone (> 90 %) is with around 100 €/kg rather

344

low45. However, when coming to a natural α-ionone, the price for even low purity product (> 86 %) is with

345

1,390 €/kg more than an order of magnitude higher46. So far, only one bulk supplier offers α-ionone

346

(> 90 %) which is labeled natural under US regulations for around 600 $/kg47. Unfortunately, no explicit

347

information could be found on the source of origin and on the enantiopurity of the product. In order to ACS Paragon Plus Environment

Page 16 of 28

Page 17 of 28

Journal of Agricultural and Food Chemistry

348

achieve economic competitiveness, the biochemical α-ionone production should be further optimized.

349

One of the bottlenecks of the process developed remains a cell density of up to 70 g-CDW/L which leads

350

to a product titer lower than elsewhere reported. As the most product was found on or in the biomass,

351

the most effort should be put into a further increase in cell density which will lead to lower specific

352

fermentation costs. The highest reported cell density of E. coli fed-batch fermentation was 148 g-CDW/L48.

353

Shiloach and Fass reviewed in their article several methods on the way to high cell density E. coli

354

fermentation. The proposed strategies were alongside the proper choice of growth medium, the intense

355

oxygen supply and the suppression of acetate accumulation during the fermentation.49

356 357

Further process optimization can be done at the interface of the fermentation and downstream process.

358

As the high concentration of α-ionone is known to be inhibitory for E. coli growth23, an in-situ product

359

recovery is an often reported approach to increase the process productivity. As we showed, most of

360

α-ionone is located in or on the cells making an ISPR quite challenging. Recovering the product from the

361

aqueous phase should indeed lead to the partitioning of more product from the biomass through the

362

fermentation supernatant to the ISPR phase. However, as the aqueous solubility of α-ionone is limited,

363

most of the product will still stick to the cells. Nevertheless, successful product removal during the

364

fermentation should immensely increase the overall process productivity. However, economics also have

365

to be taken into account. Often used two-phase fermentation requires sometimes expensive second

366

phases, e.g. n-dodecane. For n-dodecane regeneration, there will be a need for a backextraction of the

367

product with e.g. n-hexane. If the second phase is highly viscose not only the emulsification of such might

368

be challenging on an industrial scale, but also the oxygen transfer into the medium might be limited

369

decreasing the production rate. As we mentioned above, adsorption of α-ionone on a synthetic resin can

370

be easily done in-situ, or more precisely in-line. During the fermentation, a fraction of the filtered cell-free

371

supernatant can be lead through an adsorption column packed with Diaion HP20 resin. The depleted

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

372

supernatant will be fed back to the reactor. Once the adsorption column has reached its loading capacity,

373

the desorption and equilibration can be done offline. An elegant way to remove α-ionone from both the

374

solid and the liquid phase simultaneously during the production step is an in-situ stripping. Increased

375

aeration and slightly increased fermentation temperature will increase the volatility of the product leading

376

to increased partitioning to the gaseous phase. Besides, such conditions are beneficial for an increased cell

377

grows, as stated previously. After the stripping, the product can be recovered from the reactor’s off-gas

378

using absorption. Especially for an industrial scale production absorption seems to be superior compared

379

to condensation with liquid nitrogen often used on lab scale.

380 381

Increasing supplementation of products with aromas combined with the rising customer's demand for all-

382

natural formulations has led to an increased interest for the biotechnological production of aroma

383

compounds. In particular, many norisoprenoids like α-ionone are in the focus of current research. This

384

work addresses the downstream processing for a given fermentative α-ionone production process.

385

Altogether, the simple and robust overall process for the production, recovery, and purification of natural

386

enantiopure (R)-α-ionone may help to bridge the gap from the lab-scale research to the industrial scale

387

commercialization. The process developed has the potential to be transferred to the production of other

388

natural aroma compounds.

389

Acknowledgment

390

This research was supported by the Federal Ministry for Economic Affairs and Energy (BMWi) of the

391

Federal Republic of Germany (grant No. 16KN065221).

392 393

ACS Paragon Plus Environment

Page 18 of 28

Page 19 of 28

Journal of Agricultural and Food Chemistry

Nomenclature

394

𝑐𝑠𝑎𝑡

saturation aqueous solubility

[g/L]

𝑐𝑥

biomass concentration

[g-CDW/L]

𝐻𝑝𝑐

Henry’s law constant

[atm*m³/mol]

𝑙𝑜𝑔𝐾𝑂𝑊

octanol-Water partitioning coefficient

[-]

n

number of replications

[-]

𝑡

time

[min]

𝑇𝑏

boiling point

[°C]

𝑌𝑎𝑞𝑢𝑒𝑜𝑢𝑠 𝑝

specific product yield of aqueous phase

[mg/g-phase]

𝑌𝑠𝑜𝑙𝑖𝑑 𝑝

specific product yield of solid phase

[mg/g-phase]

𝑌𝑠𝑙𝑢𝑟𝑟𝑦 𝑝

specific product yield of slurry

[mg/g-phase]

ϕ

solvent phase ratio

[g-solvent/g-CWW]

Φ

desorbent phase ratio

[g-solvent/g-resin]

ψ

specific adsorbent surface ratio

[m²/ml-broth]

Ψ*

specific adsorbent surface ratio

[m²/mg-α-ionone]

395

396

Abbreviations ACN

Acetonitrile

CCD

Carotine Cleavage Dioxygenase

CDW

Cell Dry Weight

CWW

Cell Wet Weight

DAD

Diode Array Detector

F&F

Flavors and Fragrances

HPLC

High Pressure Liquid Chromatography

ISPR

In-Situ Product Recovery

ppb

Parts Per Billion

rpm rcf

Rotations Per Minute Relative Centrifugal Force

vvm

Volume per Volume per Minute

397

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442

5

References

1. Leffingwell and Associates. 2013 - 2017 Flavor & Fragrance Industry Leaders; 2018. 2. Cheetham P. Combining the Technical Push and the Business Pull for Natural Flavours. In: Berger RG, ed. Biotechnology of aroma compounds. Advances in biochemical engineering biotechnology. Berlin Heidelberg New York: Springer; 1997:1. 3. Soares M, Christen P, Pandey A, Raimbault M, Soccol CR. A novel approach for the production of natural aroma compounds using agro-industrial residue. Bioprocess Engineering. 2000;23(6):695-699. 4. Janssens L, Pooter HL de, Schamp NM, Vandamme EJ. Production of flavours by microorganisms. Process Biochemistry. 1992;27(4):195-215. 5. Berger RG. Biotechnology of flavours-the next generation. Biotechnology letters. 2009;31(11):16511659. 6. Cataldo VF, López J, Cárcamo M, Agosin E. Chemical vs. biotechnological synthesis of C13apocarotenoids: current methods, applications and perspectives. Applied microbiology and biotechnology. 2016;100(13):5703-5718. 7. Bicas JL, Silva JC, Dionísio AP, Pastore GM. Biotechnological production of bioflavors and functional sugars. Ciência e Tecnologia de Alimentos. 2010;30(1). 8. Müller DA. 2 Flavours: the Legal Framework. In: Berger RG, ed. Flavours and Fragrances. Chemistry, Bioprocessing and Sustainability: 13 Chemical Conversions of Natural Precursors. Berlin Heidelberg: Springer; 2007. 9. Larsen M, Poll L. Odour thresholds of some important aroma compounds in raspberries. Zeitschrift für Lebensmittel-Untersuchung und -Forschung. 1990;191(2):129-131. 10. Ancos dB, Ibañez E, Reglero G, Cano MP. Frozen Storage Effects on Anthocyanins and Volatile Compounds of Raspberry Fruit. J. Agric. Food Chem. 2000;48(3):873-879. 11. Robertson G. Changes in the chemical composition of volatiles released by the flowers and fruits of the red raspberry (Rubus idaeus) cultivar glen prosen. Phytochemistry. 1995;38(5):1175-1179. 12. Qian MC, Wang Y. Seasonal Variation of Volatile Composition and Odor Activity Value of‘Marion’(Rubus spp. hyb) and‘Thornless Evergreen’(R. laciniatus L.) Blackberries. Journal of Food Science. 2005;70(1):C13-C20. 13. Lalko J, Lapczynski A, Politano VT, et al. Fragrance material review on alpha-ionone. Food and chemical toxicology : an international journal published for the British Industrial Biological Research Association. 2007;45 Suppl 1:S235-40. 14. Keith ES, Powers JJ. Determination of Flavor Threshold Levels and Sub-Threshold, Additive, and Concentration Effects. J Food Science. 1968;33(2):213-218. 15. Fuganti C, Serra S, Zenoni A. Synthesis and Olfactory Evaluation of (+)- and (-)-γ-Ionone. Helvetica Chimica Acta. 2000;83(10):2761-2768. 16. O’Neil MJ, ed. The Merck index: An encyclopedia of chemicals, drugs, and biologicals. 15. ed. Cambridge: RSC Publishing Royal Society of Chemistry; 2013. 17. Lashbrooke JG, Young PR, Dockrall SJ, Vasanth K, Vivier MA. Functional characterisation of three members of the Vitis vinifera L. carotenoid cleavage dioxygenase gene family. BMC plant biology. 2013;13:156. 18. Barbosa-Filho JM, Alencar AA, Nunes XP, et al. Sources of alpha-, beta-, gamma-, delta- and epsiloncarotenes: A twentieth century review. Revista Brasileira de Farmacognosia. 2008;18(1):135-154. 19. Hiirsalmi H. The ionone content of raspberries, nectarberries and nectar raspberries and its influence on their flavour. Helsinki: Agricultural Research Centre; 1974. Annales agriculturae Fenniae. Seria Horticultura; n. 24. ACS Paragon Plus Environment

Page 20 of 28

Page 21 of 28

443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486

Journal of Agricultural and Food Chemistry

20. Vigon. Ionone Alpha Natural. [Pricing & Purchase]. 2018. Available at: https://www.vigon.com/product/ionone-alpha-natural/. Accessed September 28, 2018. 21. Jach G, Azdouffal S, Schullehner K, Welters P, Natanek A, inventors; Phytowelt Green Technologies GmbH. Method of fermentative α-Ionone production. WO2017036495A1. March 9, 2017. 22. Wang Y, inventor; ACH INNOTEK, LLC. COMPOSITIONS AND METHODS O F BIOSYNTHESIZING CAROTENOIDS AND Compositions and methods of biosynthesizing carotenoids and their derivatives. US WO 2016/154314 A1. 23. Zhang C, Chen X, Lindley ND, Too H-P. A "plug-n-play" modular metabolic system for the production of apocarotenoids. Biotechnology and bioengineering. 2018;115(1):174-183. 24. Royals EE. Cyclization of Pseudoionone by acidic reagents. Industrial & Engineering Chemistry. 1946;38(5):546-548. 25. Parliment TH. Concentration and Fractionation of Aromas on Reverse-Phase Adsorbents. Journal of Agricultural and Food Chemistry. 1981;29(4):836-841. 26. Harper M. Sorbent trapping of volatile organic compounds from air. Journal of Chromatography A. 2000;885(1-2):129-151. 27. Wardencki W, Michulec M, Curylo J. A review of theoretical and practical aspects of solid-phase microextraction in food analysis. International Journal of Food Science and Technology. 2004;39(7):703-717. 28. Reineccius G. 18 Flavour-Isolation Techniques. In: Berger RG, ed. Flavours and Fragrances. Chemistry, Bioprocessing and Sustainability: 13 Chemical Conversions of Natural Precursors. Berlin Heidelberg: Springer; 2007. 29. Lukin I, Merz J, Schembecker G. Techniques for the recovery of volatile aroma compounds from biochemical broth: A review. Flavour and Fragrance Journal. 2018;33(3):203-216. 30. Kazazi H, Rezaei K, Ghotbisharif S, Emamdjomeh Z, Yamini Y. Supercriticial fluid extraction of flavors and fragrances from Hyssopus officinalis L. cultivated in Iran. Food Chemistry. 2007;105(2):805-811. 31. Bundschuh E, Tylla M, Baumann G, Griescher K. Gewinnung von natürlichen Aromen aus Reststoffen der Lebensmittelproduktion mit Hilfe der CO2-hochdruckextraktion. Lebensmittel-Wissenschaft und Technologie;1986(19):493-496. 32. Riesenberg D, Schulz V, Knorre WA, et al. High cell density cultivation of Escherichia coli at controlled specific growth rate. Journal of Biotechnology. 1991;20(1):17-27. 33. Sander R. Compilation of Henry’s law constants (version 4.0) for water as solvent. Atmospheric Chemistry and Physics. 2015;15(8):4399-4981. 34. Griffin S, Wyllie S, Markham J. Determination of octanol–water partition coefficient for terpenoids using reversed-phase high-performance liquid chromatography. Journal of Chromatography A. 1999;864(2):221-228. 35. Etzweiler F, Senn E, Schmidt HWH. Method for measuring aqueous solubilities of organic compounds. Analytical Chemistry. 1995;67(3):655-658. 36. World Health Organization. Indoor air quality: organic pollutants: Report on a WHO Meeting Berlin 23-27 August 1987. Copenhagen: World Health Organization Regional Office for Europe; 1989; EURO SReports and Sutdies 111. 37. Sell CS. Terpenoids. In: Inc JW&S, ed. Kirk-Othmer Encyclopedia of Chemical Technology. Hoboken, NJ, USA: John Wiley & Sons, Inc; 2000. 38. Sattler KD. Thermische Trennverfahren: Grundlagen, Auslegung, Apparate. 3rd ed. Weinheim: WileyVCH; 2001.

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516

39. Bergs D, Merz J, Delp A, Joehnck M, Martin G, Schembecker G. A Standard Procedure for the Selection of Solvents for Natural Plant Extraction in the Early Stages of Process Development. Chemical Engineering & Technology. 2013;36(10):1739-1748. 40. Bergs D. A contribution to chromatographic purification of natural products. 1. Aufl. München: Verl. Dr. Hut; 2013. Schriftenreihe Anlagen- und Prozesstechnik; 7. 41. Choe E, Min DB. Mechanisms and factors for edible oil oxidation. Comprehensive Reviews in Food Science and Food Safety. 2006;5(4):169-186. 42. The European parliament and the council of the European Union. DIRECTIVE 2009/32/EC OF THE EUROPEAN PARLIAMENT AND OF THE COUNCIL of 23 April 2009 on the approximation of the laws of the Member States on extraction solvents used in the production of foodstuffs and food ingredients. 2010. Available at: https://eur-lex.europa.eu/legal-content/EN/TXT/?uri=CELEX:02009L003220100916. Accessed December 10, 2018. 43. U.S. Food & Drug Administration. Food Additive Status List. 2018. Available at: https://www.fda.gov/Food/IngredientsPackagingLabeling/FoodAdditivesIngredients/ucm091048.ht m#ftnH. Accessed December 10, 2018. 44. Sell CS. Terpenoids. In: John Wiley & Sons Inc, ed. Kirk-Othmer Encyclopedia of Chemical Technology. Hoboken, NJ, USA: John Wiley & Sons, Inc; 2000. 45. Merck KgaA. α-Ionone: ≥90%, stabilized. 2019. Available at: https://www.sigmaaldrich.com/catalog/product/aldrich/w259403?lang=de®ion=DE&cm_sp=Insit e-_-prodRecCold_xviews-_-prodRecCold5-3. Accessed 19.02.19. 46. Merck KgaA. α-Ionone: natural, ≥86%. Available at: https://www.sigmaaldrich.com/catalog/product/aldrich/w259411?lang=de®ion=DE&cm_sp=Insit e-_-prodRecCold_xviews-_-prodRecCold5-2. Accessed 19.02.19. 47. Vigon. Ionone Alpha Natural: Pricing & Purchase. 2018. Available at: https://www.vigon.com/product/ionone-alpha-natural/. Accessed September 28, 2018. 48. Korz DJ, Rinas U, Hellmuth K, Sanders EA, Deckwer W-D. Simple fed-batch technique for high cell density cultivation of Escherichia coli. Journal of Biotechnology. 1995;39(1):59-65. 49. Shiloach J, Fass R. Growing E. coli to high cell density - a historical perspective on method development. Biotechnology advances. 2005;23(5):345-357.

ACS Paragon Plus Environment

Page 22 of 28

Page 23 of 28

518 519

Journal of Agricultural and Food Chemistry

Figure Captions

520 521

Figure 1: Typical results for the fermentation using the patent pending α-ionone process delivering samples for this study.

522 523

Figure 2: Distribution of α-ionone between different phases of a fermentation broth. Mass of α-ionone extracted from separated phases and from native fermentation slurry (n-hexane,

524

ϕ = 1.0 g-solvent/g-sample, 90 minutes, one stage; n = 2).

525 526 527

Figure 3: Solvent screening for the extraction of α-ionone from the biomass (ϕ = 1.0 g-solvent/g-sample, 90 minutes; n = 2). Inlay shows the solvent ranking according to MOPASOOL© based on hydrophobicity calculations.

528 529

Figure 4: Solvent phase ratio screening. Yield of α-ionone for the extraction of wet cells with n-hexane at different solvent phase ratios for up to four stages (t = 90 min; n = 2).

530

Figure 5: HPLC chromatogram (RP C8ec 250x7 mm; H2O:ACN gradient elution) for the solvent extraction

531 532

of the biomass (n-hexane, 90 minutes, ϕ = 0.4 g-solvent/g-biomass, ambient temperature). The inlay shows an absorption spectrum scan of the peak 2.

533 534

Figure 6: Kinetics of α-ionone solvent extraction from the biomass (n-hexane, ϕ = 0.4 g-solvent/g-sample; n = 2).

535

Figure 7: Screening of the extraction temperature. Yield and purity of α-ionone for the extraction from

536

the biomass at different temperatures (n-hexane, ϕ = 0.4 g-solvent/g-biomass, 90 minutes; n = 2).

537 538

Figure 8: Adsorbent screening for the recovery of α-ionone from the aqueous supernatant of a fermentation broth (adsorption: 5 ml sample, overnight (t > 8 h); desorption: n-hexane,

539

Φ = 10 g-solvent/g-adsorber, overnight (t > 8 h); n = 2).

540

Figure 9: Screening of the desorbent phase ratio for the recovery of α-ionone from the aqueous

541 542

supernatant of a fermentation broth (adsorption: Diaion HP20, ψ = 5 m²/ml-broth, overnight (t > 8 h); desorption: n-hexane, overnight (t > 8 h) per step; n = 2).

543

TOC figure: Production and recovery of natural enantiopure (R)-α-ionone from fermentation broth.

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

figure1

figure 2

ACS Paragon Plus Environment

Page 24 of 28

Page 25 of 28

Journal of Agricultural and Food Chemistry

figure 3

figure 4

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

figure 5

figure 6

ACS Paragon Plus Environment

Page 26 of 28

Page 27 of 28

Journal of Agricultural and Food Chemistry

figure 7

figure 8

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

figure 9

TOC figure

ACS Paragon Plus Environment

Page 28 of 28