bioproduction of fumaric acid: an insight into microbial

18
REVIEW ARTICLE Bioproduction of fumaric acid: an insight into microbial strain improvement strategies Joseph Sebastian a , Krishnamoorthy Hegde a , Pratik Kumar a , Tarek Rouissi a and Satinder Kaur Brar a,b a INRS-ETE, Universit e du Qu ebec, Qu ebec, Canada; b Department of Civil Engineering, Lassonde School of Engineering, York University, Toronto, Ontario, Canada ABSTRACT Fumaric acid (FA), a metabolic intermediate, has been identified as an important carbohydrate derived platform chemical. Currently, it is commercially sourced from petrochemicals by chemical conversion. The shift to biochemical synthesis has become essential for sustainable development and for the transition to a biobased economy from a petroleum-based economy. The main limi- tation is that the concentrations of FA achieved during bioproduction are lower than that from a chemical process. Moreover, the high cost associated with bioproduction necessitates a higher yield to improve the feasibility of the process. To this effect, genetic modification of microorgan- ism can be considered as an important tool to improve FA yield. This review discusses various genetic modifications strategies that have been studied in order to improve FA production. These strategies include the development of recombinant strains of Rhizopus oryzae, Escherichia coli, Saccharomyces cerevisiae, and Torulopsis glabrata as well as their mutants. The transformed strains were able to accumulate fumaric acid at a higher concentration than the corresponding wild strains but the fumaric acid titers obtained were lower than that reported with native fumaric acid producing R. oryzae strains. Moreover, one plausible adoption of gene editing tools, such as Agrobacterium-mediated transformation (AMT), CRISPR CAS-9 and RNA interference (RNAi) mediated knockout and silencing, have been proposed in order to improve fumaric acid yield. Additionally, the introduction of the glyoxylate pathway in R. oryzae to improve fumaric acid yield as well as the biosynthesis of fumarate esters have been proposed to improve the eco- nomic feasibility of the bioprocess. The adoption of some of these genetic engineering strategies may be essential to enable the development of a feasible bioproduction process. ARTICLE HISTORY Received 23 July 2018 Revised 18 April 2019 Accepted 30 April 2019 KEYWORDS Fumaric acid; bioproduction; genetic modification; metabolic engineering; transformation; R. oryzae; fumaric acid ester Introduction Microorganisms can produce a diverse range of chemi- cals and materials. These products of microbial metab- olism, such as organic acids, have a wide range of applications and can be used in our daily life. Humans have a long history of reliance on microorganisms to produce diverse products ranging from chemicals, such as ethanol and organic acids, to consumables, such as bread, yogurt, and cheese. Fumaric acid (FA), also referred to as (E)-2-butenedioic acid or trans-1,2- ethylene dicarboxylic acid, is one such microbial meta- bolic products that have been identified as a platform chemical by the US Department of Energy (DOE) in 2004 [1,2]. It is as an important carbohydrate derived platform chemical, a key intermediate of microbial metabolism, with several applications in the synthesis of bio-based chemicals and polymer production. The diverse applications are rendered due to the presence of the dicarboxylic groups. This makes FA amenable to chemical modifications and being suitable for use as a monomer for the synthesis of biodegradable poly- mers [3]. FA is mainly used as a food additive and food acidulant and in the production of paper resins, unsat- urated polyester resins, and plasticizers [46]. The application of fumaric acid, in general, has been reviewed extensively by Das et al. [7,8] and the diverse industrial applications of this acid as a resin, plasti- cizer, and component for polymer synthesis has been reviewed by Doscher et al. [9]. More recently, deriva- tives of FA, especially fumaric acid esters (FAEs) have found its application as an important chemical with a wide array of biomedical applications, such as psoriasis and sclerosis treatment and a support material for tis- sue engineering. Clinical level studies on the pharma- cological effects of FA and its ester derivatives (FAEs) 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 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 106 CONTACT Satinder Kaur Brar [email protected], [email protected] Institut national de la recherche scientifique, Centre Eau, Terre & Environnement, 490 de la Couronne, Qu ebec (Qc) G1K 9A9, Canada ß 2019 Informa UK Limited, trading as Taylor & Francis Group CRITICAL REVIEWS IN BIOTECHNOLOGY https://doi.org/10.1080/07388551.2019.1620677

Upload: others

Post on 07-Jan-2022

3 views

Category:

Documents


0 download

TRANSCRIPT

REVIEW ARTICLE

Bioproduction of fumaric acid: an insight into microbial strainimprovement strategies

Joseph Sebastiana, Krishnamoorthy Hegdea, Pratik Kumara, Tarek Rouissia and Satinder Kaur Brara,b

aINRS-ETE, Universit�e du Qu�ebec, Qu�ebec, Canada; bDepartment of Civil Engineering, Lassonde School of Engineering, York University,Toronto, Ontario, Canada

ABSTRACTFumaric acid (FA), a metabolic intermediate, has been identified as an important carbohydratederived platform chemical. Currently, it is commercially sourced from petrochemicals by chemicalconversion. The shift to biochemical synthesis has become essential for sustainable developmentand for the transition to a biobased economy from a petroleum-based economy. The main limi-tation is that the concentrations of FA achieved during bioproduction are lower than that from achemical process. Moreover, the high cost associated with bioproduction necessitates a higheryield to improve the feasibility of the process. To this effect, genetic modification of microorgan-ism can be considered as an important tool to improve FA yield. This review discusses variousgenetic modifications strategies that have been studied in order to improve FA production.These strategies include the development of recombinant strains of Rhizopus oryzae, Escherichiacoli, Saccharomyces cerevisiae, and Torulopsis glabrata as well as their mutants. The transformedstrains were able to accumulate fumaric acid at a higher concentration than the correspondingwild strains but the fumaric acid titers obtained were lower than that reported with nativefumaric acid producing R. oryzae strains. Moreover, one plausible adoption of gene editing tools,such as Agrobacterium-mediated transformation (AMT), CRISPR CAS-9 and RNA interference(RNAi) mediated knockout and silencing, have been proposed in order to improve fumaric acidyield. Additionally, the introduction of the glyoxylate pathway in R. oryzae to improve fumaricacid yield as well as the biosynthesis of fumarate esters have been proposed to improve the eco-nomic feasibility of the bioprocess. The adoption of some of these genetic engineering strategiesmay be essential to enable the development of a feasible bioproduction process.

ARTICLE HISTORYReceived 23 July 2018Revised 18 April 2019Accepted 30 April 2019

KEYWORDSFumaric acid;bioproduction; geneticmodification; metabolicengineering; transformation;R. oryzae; fumaric acid ester

Introduction

Microorganisms can produce a diverse range of chemi-cals and materials. These products of microbial metab-olism, such as organic acids, have a wide range ofapplications and can be used in our daily life. Humanshave a long history of reliance on microorganisms toproduce diverse products ranging from chemicals,such as ethanol and organic acids, to consumables,such as bread, yogurt, and cheese. Fumaric acid (FA),also referred to as (E)-2-butenedioic acid or trans-1,2-ethylene dicarboxylic acid, is one such microbial meta-bolic products that have been identified as a platformchemical by the US Department of Energy (DOE) in2004 [1,2]. It is as an important carbohydrate derivedplatform chemical, a key intermediate of microbialmetabolism, with several applications in the synthesisof bio-based chemicals and polymer production. Thediverse applications are rendered due to the presence

of the dicarboxylic groups. This makes FA amenable tochemical modifications and being suitable for use asa monomer for the synthesis of biodegradable poly-mers [3].

FA is mainly used as a food additive and foodacidulant and in the production of paper resins, unsat-urated polyester resins, and plasticizers [4–6]. Theapplication of fumaric acid, in general, has beenreviewed extensively by Das et al. [7,8] and the diverseindustrial applications of this acid as a resin, plasti-cizer, and component for polymer synthesis has beenreviewed by Doscher et al. [9]. More recently, deriva-tives of FA, especially fumaric acid esters (FAEs) havefound its application as an important chemical with awide array of biomedical applications, such as psoriasisand sclerosis treatment and a support material for tis-sue engineering. Clinical level studies on the pharma-cological effects of FA and its ester derivatives (FAEs)

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

61

62

63

64

65

66

67

68

69

70

71

72

73

74

75

76

77

78

79

80

81

82

83

84

85

86

87

88

89

90

91

92

93

94

95

96

97

98

99

100

101

102

103

104

105

106

CONTACT Satinder Kaur Brar [email protected], [email protected] Institut national de la recherche scientifique, Centre Eau,Terre & Environnement, 490 de la Couronne, Qu�ebec (Qc) G1K 9A9, Canada� 2019 Informa UK Limited, trading as Taylor & Francis Group

CRITICAL REVIEWS IN BIOTECHNOLOGYhttps://doi.org/10.1080/07388551.2019.1620677

have confirmed their efficacy. Methyl esters of fumaricacid, mono-methyl fumarate (MMF) and dimethylfumarate (DMF) esters, has been even licensed for usein Europe for the treatment of severe psoriasis vulgaris[10,11]. The recent approval of FDA of DMF (Tecfidera)for use in the treatment of relapsing forms of multiplesclerosis, proves the importance of FA and FAEs. Inaddition, FAEs, specifically their ethyl esters, has foundits application in tissue engineering as a scaffold aswell as for drug delivery [12,13]. The review by Daset al. [8] provides comprehensive information aboutthe diverse medical applications of fumaric acid andits derivatives.

The diverse applications of FA have fueled thedemand for dicarboxylic acid, which is fulfilled by themeans of chemical synthesis from petroleum prod-ucts. Emphasis on the transition towards the bio-based economy has led to the recognition of interestin the development of bioproduction strategies forFA production. FA is a key intermediate of the meta-bolic pathway of a diverse group of microorganismsand has led to the investigation of these microorgan-isms as a potential means of FA production. Fungalstrains such as Rhizopus oryzae, has been identified asa FA overproducer and a potential source for FA. Thisreview discusses various strain improvement strat-egies to improve R. oryzae mediated FA productionas well as metabolic and genetic engineering strat-egies investigated to achieve microbial productionusing non-FA producers. These investigations havenot led to commercial bioproduction of FA andthe plausible reasons for this have been provided.Moreover, the advancement in metabolic engineeringtechniques have provided us with novel tools andpossible avenues for the use of these strategies forthe bioproduction of FA and its derivatives. Previousreviews have emphasized primarily on the control ofprocess parameters in order to improve and controlFA production. These reviews have provided little dis-cussion on genetic engineering strategies that can beadopted to improve FA production. This lack of com-piled information on strain improvement strategies,especially genetic engineering, is attempted to beconsidered in this review.

Current status of the commercial synthesis offumaric acid

At present, the global demand for FA is solely fulfilledby petrochemical methods of synthesis as representedin Figure 1. It is derived by the isomerization of maleicacid obtained from hydrolysis of maleic anhydride.

The maleic anhydride is obtained by the oxidation ofbutane or benzene in the presence of the catalyst,vanadyl pyrophosphate [3,4]. The catalyst mediatedconversion is the only feasible technique for theselective conversion of butane into maleic acid [14].This chemical catalyst mediated process provides ahigh process yield of FA (112% w/w) from maleicanhydride [3,14]. However, this process also leads tothe formation of the toxic gas, carbon monoxide, andgreenhouse gas, carbon dioxide, as byproducts, whichin turn leads to environmental pollution and contrib-utes to global warming [7]. The global demand for FAis expected to expand to over 300 Kilotons by theyear 2020 from 225.2 Kilotons (2012). This increase indemand is mainly fueled by a rise in the demand forthe acid used in the food and beverage industry,which accounts for over 35% of global consump-tion [15–17].

Moreover, FA is an important intermediate in thesynthesis of edible products, such as L-malic acid andL-aspartic acid. The increase in demand for artificialsweeteners, such as aspartame, in beverages andfoods, will further increase the worldwide demandfor FA and its derivatives [18]. Furthermore, therecent identification of its pharmaceutical propertiescan lead to increased usage of FA and its derivativesin the pharmaceutical industry, as discussed earlier.This will further enhance the global demand for thisacid. Additionally, the high price of crude oil, itsdepletion and the recent emphasis on green chemis-try for sustainable development, has necessitatedthe adoption of alternate means for increased FAproduction [15,19].

Bioproduction of fumaric acid

Microbial bioproduction of this carbohydrate derivedbiochemical provides an attractive alternative whichensures sustainable development by reducing thedependence on fossil fuels. Moreover, the use ofrenewable feedstocks helps the recovery of valuablecarbon. FA is one such compound synthesized bymany microorganisms, especially aerobic microbes, asan intermediate compound during the TCA cycle[7,20]. As early as 1911 the fungal strain, Rhizopus nig-ricans, was identified for its ability to synthesize FA asearly as 1911. Later studies, involving multiple fungalspecies, concluded that Rhizopus species was the bestproducer of fumaric acid. Xu et al. [5] provided a briefsummary of the history of the bioproduction of FA.Additionally, the adoption of a bioproduction strategy

107

108

109

110

111

112

113

114

115

116

117

118

119

120

121

122

123

124

125

126

127

128

129

130

131

132

133

134

135

136

137

138

139

140

141

142

143

144

145

146

147

148

149

150

151

152

153

154

155

156

157

158

159

160

161

162

163

164

165

166

167

168

169

170

171

172

173

174

175

176

177

178

179

180

181

182

183

184

185

186

187

188

189

190

191

192

193

194

195

196

197

198

199

200

201

202

203

204

205

206

207

208

209

210

211

212

2 J. SEBASTIAN ET AL.

allows for carbon dioxide fixation and recovery, espe-cially when organic industrial residues are employed.

However, the advent of less expensive petrochem-ical means of fumaric acid synthesis led to the discon-tinuation of the bioproduction of FA. Higher yields ofFA are obtained during the chemical process (112%w/w against 85% w/w via fermentation using glucoseas carbon source). Also, the lower production cost andbetter control over production conditions led to thewidespread use of the chemical process for FA pro-duction [3]. The recent increase in the price of petrol-eum products, increasing the emphasis on low carbonfootprint production strategies and green chemistryapproaches for sustainable development. This hasresulted in the resurgence of interest in fumaric acidproduction by fermentation [5,15]. Moreover, theadoption of bioproduction strategy allows for the util-ization of renewable biomass as well as industrialresidues for FA production. This in turn significantlyreduces the production costs thereby improvingits feasibility. Additionally, the low solubility of FA,5–7 g/L (at room temperature), ensures its easy

recovery by simple precipitation following acidification.Hence, no specialized strategies are required for therecovery of FA [21,22]. The FA crystals so recoveredhave been observed to be of high purity, with inor-ganic components in the fermentation medium beingthe primary contaminant, and can be used thereafterfor low purity applications such as the manufacture ofresins or polymers as well as cattle feed supplements[3]. To further improve FA purity, the use of activatedcarbon has been investigated and found to be effect-ive [23,24]. The recovered FA (of purity >99%) can beused for biomedical applications such as psoriasisand multiple sclerosis treatment.

Metabolic pathway of fumaric acid synthesis

Fumaric acid is an important intermediate of the tri-carboxylic acid (TCA) cycle (Krebs cycle). Two differentmetabolic pathways are mainly involved in the synthe-sis of FA viz. oxidative TCA cycle and reductive TCApathway. Fixation of CO2 and conversion of pyruvateto oxaloacetate occurs during the TCA cycle. Reductive

213

214

215

216

217

218

219

220

221

222

223

224

225

226

227

228

229

230

231

232

233

234

235

236

237

238

239

240

241

242

243

244

245

246

247

248

249

250

251

252

253

254

255

256

257

258

259

260

261

262

263

264

265

266

267

268

269

270

271

272

273

274

275

276

277

278

279

280

281

282

283

284

285

286

287

288

289

290

291

292

293

294

295

296

297

298

299

300

301

302

303

304

305

306

307

308

309

310

311

312

313

314

315

316

317

318

Figure 1. Chemical fumaric acid synthesis and its derivatives.

CRITICAL REVIEWS IN BIOTECHNOLOGY 3

CO2 fixation catalyzed by the pyruvate carboxylaseenzyme is the main reason for the accumulation offumaric acid during fermentation as it was observed inthe case of the Rhizopus species [5]. The reductive cyclehas a maximal theoretical yield of 2 moles of FA permole of glucose consumed and 2 moles of CO2 fixed.Reliance on reductive pyruvate carboxylation, as thesole pathway, will lead to no ATP formation. This ATP isrequired for cell maintenance and will affect overall cellgrowth. Hence, the TCA cycle is active during FA pro-duction and complete theoretical fumaric yield willnot be achieved [3,7]. Adoption of strategies to limitfungal growth has the potential to ensure optimumfumaric acid production. The metabolic pathway andenzymes involved in the key steps of this pathway arerepresented in Figure 2 and a brief summary of themetabolic pathways involved in the microbial fumaricacid production is provided in reviews by Engel et al.[3] and Xu et al. [5]. Additionally, it was identifiedthat amino acid and fatty acid metabolism, as wellas activation of the glyoxylate pathway, can have apotential impact on FA accumulation [25–27].

Rhizopus sp. mediated fumaric acid production

Fungal species have been employed to produce a vastarray of compounds and consumables. Multicellular

filamentous fungi or molds are extensively used toproduce fermented foods, secondary metabolites, andindustrial enzymes. Rhizopus sp. is an example of fila-mentous fungi, belonging to the “Mucoraceae” family,that is used to produce a wide array compoundsranging from fermented food products to platformchemicals such as fumaric acid [20,28,29].

Rhizopus oryzae, generally regarded as safe (GRAS),has usually been employed for FA production via fer-mentation at neutral pH values to produce fumaratesalts [22]. This fungal species is subdivided into twogroups (Type I and Type II) according to their geneticand phenotypic differences. Type I accumulates lacticacid whereas Type II mostly synthesize FA in the pres-ence of fermentable sugars. The main differencebetween these two types is that Type I contains twolactate dehydrogenase (ldh) genes, ldhA, and ldhB,whereas Type II only has one of the genes ldhB. Thelactate dehydrogenase enzyme, LdhA, encoded by theldhA gene enables the conversion of pyruvic acid tolactic acid whereas the enzyme encoded by ldhB genepossesses insufficient reductive capacity. This reducedcapacity allows for the conversion of excess pyruvateto fumaric acid and/or ethanol [28,30]. Table 1 pro-vides a summary of FA production studies conductedusing R. oryzae.

319

320

321

322

323

324

325

326

327

328

329

330

331

332

333

334

335

336

337

338

339

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

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

Figure 2. Overview of metabolic pathway and enzymes involved in the fumaric acid synthesis in R. oryzae. EMP stands forEmbden–Meyerhof–Parnas pathway or glycolysis. The enzymes involved in the various steps of the pathway are PDC—pyruvatedecarboxylase (EC 4.1.1.1); ADH—alcohol dehydrogenase (EC 1.1.1.1); LDH—lactate dehydrogenase (EC 1.1.1.27); PDH—PYRUVATEdehydrogenase (EC 1.2.4.1); PYC—pyruvate carboxylase (EC 6.4.1.1); MDH—malate dehydrogenase (EC 1.1.1.37) and FUM—fumarase (EC 4.2.1.2). For color version of the figure refer e-print.

4 J. SEBASTIAN ET AL.

Limitations of R. oryzae based production

R. oryzae mediated FA production is usually per-formed at neutral pH values. The production of theacid during fermentation leads to a significantdecrease in pH from 6 to 4 within the first 48 h offermentation. This reduction in pH causes the rate ofproduction to slow down and eventually cease com-pletely. To prevent this drop in pH, the fumaric acidis converted to fumarate salts by using neutralizers,such as calcium carbonate and sodium carbonate.Fumaric acid is then recovered from the fermentationbroth by precipitation, due to its low solubility, fol-lowing acidification. The requirement of neutralizersand acids during the recovery step leads to highconsumption of bases and acids as along with theproduction of large quantities of salts. Calcium car-bonate is commonly used as the neutralizer butthe insolubility of calcium salts causes considerablepower consumption and process difficulties, such asincreased viscosity and associated power consump-tion [22,39,48]. Hence, sodium carbonate is mainlyproposed as a neutralizing agent because the sodiumsalt of fumarate is soluble. Gangl et al. [48] have per-formed a comprehensive comparison of the advan-tages and disadvantages of using calcium carbonateand sodium carbonate as the neutralizing agent.

Some of the above-mentioned limitations of thefungal strain mediated fumaric acid fermentation canbe overcome by performing the fermentation at alower pH. The move to a lower pH fermentationreduces the quantity of neutralizing agent required inorder to maintain pH close to neutral as well as thequantity of salt generated during the recovery offumaric acid from the fermented broth. Moreover, the

undissociated fumaric acid so formed can be recov-ered by the crystallization from the broth due to thelow solubility of FA. Additionally, the direct recoveryof fumaric acid from the broth using adsorbents, suchas polyvinyl pyridine (PVP) resin and Amberlite IRA-900, would be possible [5]. This strategy of FA recov-ery using adsorbents was successfully used by Caoet al. [47] to recover high purity FA following desorp-tion and acidification. To this effect, the shortenedduration of pH control and fermentation at pH 5 wasproposed [22]. It was observed that eliminating off pHcontrol at 90 h instead of 120 h did not lead toimportant effects on FA production. The drawback ofthis strategy is that pH control is required during theinitial stages of fermentation even though reducedconsumption of the neutralizing agent is achieved.

Moreover, it has been observed that the morph-ology of the fungus during fermentation has a pro-found impact on FA production and has beenextensively studied by Engel et al. [22] and Zhou et al[49]. It was also observed by them that the fungalpellet diameter played a key role in FA productivity.Zhou et al. [49] concluded that lower pellet diameterenhanced for mass transfer as well as better contactwith more of the actively growing fungi presenton the surface of the culture. Furthermore, it wasobserved by Engel et al. [22] that an anaerobic envir-onment was present at the core of the fungal pellet.Therefore, fungal pellets of smaller diameter are pre-ferred which requires the optimization of conditions,such as agitation rate, pH conditions, temperature,and working volume, to achieve the optimum pelletdiameter. This dependence on fungal morphologyfurther adds to the complexities associated with

425

426

427

428

429

430

431

432

433

434

435

436

437

438

439

440

441

442

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

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

517

518

519

520

521

522

523

524

525

526

527

528

529

530

Table 1. Summary of fumaric acid production studies carried out using R. oryzae.

Carbon source FermenterFumaric acidconcentration

Productivity(g/L/h) Reference

Glucose Stirred tank 56.2 g/L 0.7 [31]Glucose Stirred tank 41.1 g/L 0.37 [32]Glucose Stirred tank 32.1 g/L 0.32 [33]Glucose Stirred tank 30.2 g/L 0.19 [22]Xylose Shake flask 28.4 g/L – [34]Xylose Shake flask (immobilization) 45.3 g/L – [35]Glucose/xylose Shake flask 46.7 g/L – [36]Corn straw Shake Flask 27.8 g/L 0.33 [37]Cornstarch Shake flask (Simultaneous saccharification fermentation) 44.1 g/L 0.53 [38]Cornstarch Shake Flask 45.0 g/L 0.55 [32]Brewery wastewater Shake Flask 31.3 g/L – [39]Apple juice extraction waste Shake Flask 25.2 g/L 0.35 [40]Diary manure Stirred tank 31.0 g/L 0.32 [41]Apple pomace Solid state-Rotating drum fermenter 138 g/kg dry weight – [40]Glucose/crude glycerol Shake flask 22.81 g/L 0.34 [42]Synthetic medium Immobilized fungi 32.03 g/L 1.33 [43]Synthetic medium Immobilized fungi 40.13 g/L 0.32 [44]Synthetic medium Immobilized fungi 30.3 g/L 0.21 [45]Brewery wastewater Immobilized fungi 43.67 g/L 1.21 [46]Synthetic medium Immobilized fungi- fed-batch 85 g/L 4.25 [47]

CRITICAL REVIEWS IN BIOTECHNOLOGY 5

fungal bioproduction of fumaric acid. To this effect,immobilization of fungi on the support have beenproposed to reduce the effects of pellet morphologyand has been found effective in increasing FA prod-uctivity as well as its yield [43,46,47]. Reduction in theduration of fermentation, leading to increased prod-uctivity, from 144 h to 24 h was observed by Gu et al.[43]. The fumaric acid yield of 32.03 g/L was obtainedpost immobilization of the fungi R. arrhizus duringthis study. In addition, a glucose to nitrogen ratio ofthe production media has been observed to have aprofound impact on FA production. To enable max-imum FA yield, nitrogen limitation is recommended tolimit fungal growth and achieve optimal diversion ofsugars towards fumarate production. It has also beenobserved that optimal activity of the cytosolic fumar-ase enzyme was obtained under nitrogen limitingconditions and this activity increased 300% uponreduction of the urea concentration from 2.0 to0.1 g l–1 [3,19,50].

To overcome the above-mentioned complexitiesof R. oryzae mediated fumaric acid production, thegenetic transformation of microorganisms, such asS. cerevisiae (capable of growing at low pH), to accu-mulate fumarate has been considered. These strat-egies of genetic engineering microorganisms toproduce fumaric acid are discussed in the follow-ing sections.

Improved fumaric acid production throughstrain improvement and genetic modification

The lower yield of FA, during bioproduction, com-pared to chemical synthesis has motivated researchersto investigate strategies for improving the FA yieldof the filamentous fungus R. oryzae. These strainimprovement strategies primarily rely on increasingproduct yields thereby enhancing the feasibility ofcommercial bioproduction. These techniques of strainimprovement have been extensively reviewed byMeussen et al. [51] and Li et al. [52] but strainimprovement techniques targeted towards FA

production have so far not been discussed and thisdiscussion occurs in later sections of this review.Moreover, under low pH conditions (3–5), due toreduced fungal growth as well as the complexities ofthis fermentative fungal strain mediated fumaric acidproduction has led researchers to investigate the pos-sibility of using alternate microorganisms for the pro-duction of FA. Developments in genetic engineeringtechniques, such as genetic transformation and over-expression of homologous genes, such as fum, pycand mdh, and metabolic engineering, have providedthe tools for improving microbial production oforganic acids, such as: malic acid, succinic acid, andfumaric acid [29,51].

Strain improvement strategies for R. oryzae

Random mutagenesis

Random mutagenesis involves the use of chemicalmutagens, such as diethyl sulfate, nitrosoguanidineand diethyl sulfate, or radiation, such as UV light andgamma radiation, to induce changes in the geneticmakeup of an organism. This technique has for centu-ries been employed for the development of novelstrains. The main drawback of this technique is thatfollowing induction of the mutation, multiple mutantstrains are generated. These strains generate the needto be screened and a suitable strain selected which istime-consuming and laborious [19,51]. The mutantstrains selected were able to produce higher yieldsof fumaric acid than the parent strain but these titerswere significantly lower than the maximum titersreported for strains obtained from culture collections(represented in Table 2) but comparable to other titersthat have been reported [31–33]. Hence, randommutagenesis can be used as a viable tool for strainimprovement. The adoption of novel strategies, dis-cussed in a later section, in combination with randommutagenesis can be useful for the selection of appro-priate high yielding strains. Q1Moreover, the use ofthis technique generates multiple mutations in the

531

532

533

534

535

536

537

538

539

540

541

542

543

544

545

546

547

548

549

550

551

552

553

554

555

556

557

558

559

560

561

562

563

564

565

566

567

568

569

570

571

572

573

574

575

576

577

578

579

580

581

582

583

584

585

586

587

588

589

590

591

592

593

594

595

596

597

598

599

600

601

602

603

604

605

606

607

608

609

610

611

612

613

614

615

616

617

618

619

620

621

622

623

624

625

626

627

628

629

630

631

632

633

634

635

636

Table 2. Random mutagenesis of R. oryzae and fumaric acid concentrations obtained (R. oryzae reassigned as R.arrhizus, [53]).Fungal strain Mutagen Fumaric acid titer (g/L) Yield (g/g) Reference

Pure cultureR. arrhizus NRRL 2582 – 103 0.79 [54]R. arrhizus NRRL1526 – 98 0.82 [55]

MutantR. oryzae RUR709 UV and c-ray mutagenesis 32.1 0.45 [33]R. oryzae ZD-35 UV irradiation 57.4 0.67 [32]R. oryzae ME-F01 UV and nitrosoguanidine 52.7 – [56]R. oryzae ME-F12 Nitrogen ion implantation 44.1 0.44 [38]R. oryzae FM19 Femtosecond laser irradiation 49.4 0.56 [25]

6 J. SEBASTIAN ET AL.

Author Query
AQ1: Please note that the section number 4.1.4 has been deleted as there is no such section number under the main Section 4.

microorganism and the other metabolic activities maybe hampered in addition to the targeted metaboliteproduction. Also, the damage to the DNA by the useof high energy radiations may prevent DNA replicationand hamper the overall transfer of the desired trait.

Homologous and heterologous gene expression inR. oryzae for fumaric acid production

Fumaric acid is an intermediate of the TCA cycle, asmentioned in the earlier section, but the fumaricacid generated during the oxidative pathway is notaccumulated as it is consumed for cell growth andmaintenance. The accumulation of fumaric acidoccurs via the reductive TCA pathway, as observedin the case of R. oryzae. Multiple enzymes areinvolved in the synthesis of fumaric acid from pyru-vate. Pyruvate carboxylase (PYC) catalyzes the ATP-dependent condensation of pyruvate and CO2 toform oxaloacetic acid (OAA). The OAA, so formed,is converted to malate by malate dehydrogenase(MDH), followed by conversion of malate to fumarateby the enzyme fumarase. [57] The overexpression ofthe enzymes involved should be able to increasefumarate production.

Zhang et al. [57] observed a 26% increase infumaric acid production in their study by transformingR. oryzae to express homologous and heterologousenzymes. They introduced the exogenous phosphoe-nolpyruvate carboxylase enzyme (PEPC) in addition toPYC using the plasmid (pPyrF2.1A) mediated trans-formation of fungi R. oryzae M16. The PEPC enzyme,not present in R. oryzae, is involved in the conversionof phosphoenolpyruvate to OAA with CO2 fixation.Increased PYC and PEPC activity of 3–6mU/mg wereobserved. The redirection of this metabolic pathwayresulted in the introduction of these genes that leadto an increase in fumarate production [57]. In anotherstudy, the impact of the overexpression of the fumRgene that encodes for the fumarase enzyme wasinvestigated. It was observed that the over-expressionof fumarase in R. oryzae lead to malic acid productioninstead of FA. The over-expressed fumarase is alsoable to catalyze the hydration of FA to malic acid.Hence, it can be concluded that the presence offumarase alone may not be responsible for FA accu-mulation during R. oryzae mediated fermentation [58].To this effect, the observation by Peleg et al. has aparticular significance [59]. They identified two isoen-zymes of fumarase with a different localization withinthe cell. One isoenzyme was solely present in thecytosol whereas the other was found in both thecytosol and in the mitochondria. It was also observed

that the inhibition of fumarase activity by cyclohexi-mide only affected fumaric acid production anddecreased the activity of the cytosolic fumarase isoen-zyme [59]. Therefore, the transformation of R. oryzaewith the appropriate isoenzyme has the potential toimprove FA production as the reductive TCA cycleoccurs in the cytosol.

In addition, a dicarboxylic acid transporter withhigh selectivity for fumaric acid can also play animportant role in its production with R. oryzae. Thisconclusion is supported by the observation that malicacid production by Aspergillus oryzae NRRL 3488 wasimproved twofold by the overexpression of a C4-dicar-boxylate transporter. The overexpression of malatedehydrogenase and pyruvate carboxylase in conjunc-tion with the transporter led to an additional increaseof 27% and achieved a malic acid concentration of154 g/L in 164 h [60]. The high concentration of malicacid achieved in this study points to the possibility ofachieving a similar titer for fumaric acid by providingan efficient system for the conversion of the malicacid into fumaric acid.

Fumaric acid production in recombinantmicroorganisms

The limitations of R. oryzae mediated synthesis offumaric acid, discussed previously in Limitations of R.oryzae based production section, has motivatedresearchers to investigate the possibility of usingother microorganisms, particularly Escherichia coli andSaccharomyces cerevisiae, to produce fumaric acid. Tothis effect, the metabolic pathways that influence FAproduction have to be investigated prior to the gen-eration of recombinant strains. Cellular metabolismis highly regulated and made up of a plethora of vari-ous components, such as transcripts, proteins, andmetabolites, and the production of the compound ofinterest is directly linked to the cellular activities.Hence, an accurate understanding of the metabolicstates and regulatory mechanisms are important.

The most effective way to evaluate the metabolicstate is to quantify various metabolites. Therefore, theadvances in analytical techniques, such as NMR, gaschromatography-mass spectrometry (GC-MS), gaschromatography time-of-flight mass spectrometry (GC-TOF) and liquid chromatography-mass spectrometry(LC-MS), has enabled high-throughput analysis of themetabolites. This has allowed for quick comparativeanalysis of metabolite profiles post-genetic modifica-tion and provides an understanding of the physio-logical state of the cell, which in turn provides us

637

638

639

640

641

642

643

644

645

646

647

648

649

650

651

652

653

654

655

656

657

658

659

660

661

662

663

664

665

666

667

668

669

670

671

672

673

674

675

676

677

678

679

680

681

682

683

684

685

686

687

688

689

690

691

692

693

694

695

696

697

698

699

700

701

702

703

704

705

706

707

708

709

710

711

712

713

714

715

716

717

718

719

720

721

722

723

724

725

726

727

728

729

730

731

732

733

734

735

736

737

738

739

740

741

742

CRITICAL REVIEWS IN BIOTECHNOLOGY 7

with an understanding of their metabolic status. Tofurther improve an understanding of the metabolicstate of the cell, isotopomer experiments can be per-formed. These studies provide us with additionalinformation on intracellular fluxes and metabolic rates.13C-labeled glucose is the most frequently used iso-topomer. The monitoring of the distribution of theisotopomer, as the labeled glucose is metabolized,can be used to estimate metabolic reaction rates.Hence, provide quantitative information regarding car-bon flow through the metabolic pathways [60,61]. Anunderstanding of this would allow for the selection ofthe metabolic pathway of interest that would enablemaximum conversion of glucose to fumaric acid withreduced byproduct formation.

Improvement of the cellular systems by metabolicengineering strategies is vital to enhance the overallproductivity and yield of the target compound. Thedevelopment of stoichiometric metabolic modelswhich includes all enzymatic reactions of organisms,such as E. coli and S. cerevisiae, has enabled the pre-diction of metabolic states using in silico simulations.Metabolic flux analysis (MFA), flux balance analysis(FBA), and metabolic pathway analysis (MPA) are themost popular tools in stoichiometric metabolic net-work analysis. Additionally, database resources, suchas UniProt, BRENDA, BioCyc, KEGG, and the EnzymeCommission database, can be used for the reconstruc-tion of a metabolic pathway [29,62]. The above-men-tioned tools for metabolic engineering have beenextensively used for transformation of E. coli and S.cerevisiae strains but have not been readily used forthe transformation of R. oryzae. There are only a fewreports the transformation of R. oryzae that adoptedmetabolic flux analysis for metabolic engineering ofconsidering this fungal strain.

One such early study used TELUX, a specific radio-activity curve-matching program, to investigate glu-cose metabolism by R. oryzae. In the study, 14Cisoprotomer labeled glucose was used to understandglucose metabolism and thereby increase lactate yield.This was followed by UV-mediated random mutagen-esis and the selection of a lactate overproducingstrain. The prediction of metabolic pathway modifica-tion that led to lactate overproduction, provided byglucose flux analysis, helped in the identification of asuitable strain. This selected strain showed reducedethanol production and chitin synthesis, as predicted,and a lactic acid yield of 30 g/L was obtained [63]. Inanother study, the results of the carbon flux analysisof a mutant lactic acid overproducing strain, R. oryzaeR1021, and the parent strain, R. oryzae R3017, showed

that pyruvate was the key metabolic intermediate. Themutant strain showed reduced loss of carbon to etha-nol and acetyl-CoA and provided a lactic acid yield of79.4 g/L, which was 52% higher than that of the par-ent strain [64]. These studies focused on the synthesisof lactic acid but the understandings can possibly beadopted for FA production. Such a study on metabolicprofiling, post femtosecond laser-mediated randommutagenesis, identified that modifications to aminoacid and fatty acid metabolism led to a 1.59-foldincrease in FA production [25]. These studies usedMFA or metabolic profiling for understanding carbonmetabolism either before or after random mutagenesisbut not for designing metabolic pathways and trans-forming the fungi. The importance of a better under-standing of metabolic activities using metaboliteprofiling is further signified by the observation thatthe FA production was enhanced by 39.7% as a resultof supplementation with 1% linoleic acid during R.oryzae mediated fermentation at low pH (pH 3) condi-tion. The strategy of supplementation with linoleicacid was adopted based on the understanding of thechanges to the metabolic activity at low pH conditionsprovided by GC-MS assisted metabolite profiling [65].The use of in silico metabolic modeling was found tobe effective in designing a metabolic engineeringstrategy to achieve a higher FA titer in T. glabrata[66]. Hence, an in silico model of fumarate synthesisof R. oryzae is essential to achieve effective conversionof glucose to fumarate.

The successful adoption of MFA assisted strainselection, in combination with strain improvementstrategies, can also be adopted to improve FA yield inR. oryzae. MFA can act as an important tool for theselection of an appropriate high FA producing strainfollowing random mutagenesis. Moreover, it would bepossible to obtain far higher titers for FA by usingMFA mediated design and then transform the fungalstrain using targeted genetic modification approaches.These strategies would enable reduced byproductsynthesis, such as ethanol and malic acid, as wellas effective redirection of a metabolic pathway thatallows for maximum productivity by optimal conver-sion of glucose to FA, as observed in case of both E.coli and S. cerevisiae discussed below.

Escherichia coli

High concentrations of fumaric acid are not normallyproduced in E. coli and 20 g/L has been shown tohave an impact on cell growth [26]. The adoption ofgenetic engineering strategies has allowed the useof E. coli for the production of a diverse range of

743

744

745

746

747

748

749

750

751

752

753

754

755

756

757

758

759

760

761

762

763

764

765

766

767

768

769

770

771

772

773

774

775

776

777

778

779

780

781

782

783

784

785

786

787

788

789

790

791

792

793

794

795

796

797

798

799

800

801

802

803

804

805

806

807

808

809

810

811

812

813

814

815

816

817

818

819

820

821

822

823

824

825

826

827

828

829

830

831

832

833

834

835

836

837

838

839

840

841

842

843

844

845

846

847

848

8 J. SEBASTIAN ET AL.

compounds and possibly fumarate. This possibility hasbeen investigated. The de-repression of the glyoxylatepathway in combination with overexpression anddeletion of specific genes was exploited to induce FAaccumulation [26,27]. Song et al. [26] used rationalmetabolic engineering strategies coupled with in silicoflux response simulation to develop a novel E. colistrain. The in silico flux modeling simulation (E. coligenome-scale metabolic model EcoMBEL979) wasused to investigate the influence of PEPC flux onfumarate production and helps to understand theresponse to a metabolic shift. It was observed theoverexpression of PEPC was able to increase the FAproduction and further genetic modifications selectedbased on their effect on FA production. The final con-centration of FA was achieved by transforming E. colito overproduce FA via the noncyclic glyoxylate route.The isocitrate lyase repressor gene iclR was deletedto enhance the carbon flow into the glyoxylate shuntin addition to the silencing of fumarase isozymes.Moreover, endogenous PEPC was also overexpressedto increase the reductive TCA cycle flux to achieve FAproduction. Additionally, genes involved in the furtherconversion of FA and byproduct syntheses, such asarcA, aspA, lacL, and ptsG, were deleted. FA produc-tion with the final strain obtained was 28.2 g/L, and aproductivity of 0.448 g/L/h under aerobic fed-batchculture conditions [26].

The glyoxylate pathway flux was enhanced simi-larly to obtain a genetically modified strain of E. colicapable of utilizing glycerol as a feedstock for FAproduction. The conversion of fumaric acid to malicacid was prevented by the deletion of fumA, fumB,and fumC genes that code for three fumarases. Itwas observed that the overexpression of glyoxylateshunt and PEPC enzyme led to enhanced fumaricacid production as well as reduction of acetate for-mation (a byproduct). Additionally, the gene codingfor aspartase, which converts fumarate to aspartate,was deleted to prevent the reverse conversion. AnFA concentration of 41.5 g/L was achieved with theoverall productivity of 0.51 g/L/h following fermenta-tion under fed-batch conditions. This correspondedto 70% of the maximum theoretical yield [27]. Theconcentration of FA achieved during these investiga-tions is still lower than the titers obtained by cultur-ing R. oryzae. These studies focused primarily on theredirection of the TCA cycle towards the glyoxylatepathway for FA accumulation. However, the reduc-tive TCA pathway has the potential to produce 2moles of FA for every mole of glucose. This yield isdouble than that of the theoretical maximum of the

oxidative glyoxylate pathway (1mol/mol glucose)[67]. Hence, it may be possible to obtain even ahigher titer for FA with the enhancement of thereductive TCA pathway along with the genetic modi-fication strategies adopted during these studies.

Another metabolic pathway that has been investi-gated for fumarate bioproduction in E. coli was theintroduction of the mammalian urea cycle [68].Fumaric acid is produced as a byproduct in this meta-bolic pathway during the conversion of L-arginosucci-nate to arginine. The recombinant E. coli (ABCDIA-RAC) was produced by blocking the Krebs cycle andintroducing the urea cycle. To improve FA production,several genes, such as fumA, fumB, fumC, frdABCD,iclR, and arcA, were deleted. Additionally, the ureacycle was introduced by overexpressing the homolo-gous carAB, argI, and heterologous rocF genes whichencoded for carbamoyl phosphate synthase (CPS),ornithine carbamoyl transferase (OTC), and arginase,respectively. The strain isolated was able to yield afumarate titer of 11.38mmol/L (1.32 g/L) and more-over, it was concluded that cell growth and the dere-pression of the Krebs cycle were required for high-titer FA production in aerobic conditions [68]. Inanother study by the same research group, the FAyield of the strain was further improved to 9.42 g/L bythe overexpression of the homologous C-4 dicarboxy-late carrier gene dcuB [69]. These studies have failedto achieve the FA yields reported for R. oryzae butshows that targeted multilevel genetic transformationsare essential to improve fumarate yield.

Saccharomyces cerevisiae

The yeast, S. cerevisiae is a well-established industrialproduction organism and it possesses exceptional cul-tivation characteristics. They require only simplemedia for growth and are tolerant to the presence ofinhibitors. Moreover, there is a high degree of toler-ance to (high) sugar concentrations and their abilityto grow under acidic conditions. The inability to growunder low pH conditions is one of the main limita-tions to R. oryzae mediated FA production. S. cerevi-siae is not naturally capable of accumulating FA andthe feasibility of the induction of FA accumulation byengineering S. cerevisiae has been investigated [5].

The introduction of heterologous cytosolic R. oryzaemalate dehydrogenase, high-level expression of the R.oryzae fumarase and overexpression of the nativecytosolic pyruvate carboxylase was observed to onlyyield an FA concentration of 3.2 g/L [70]. In a similarstudy, simultaneous co-expression of both reductiveand oxidative routes of FA was investigated. This

849

850

851

852

853

854

855

856

857

858

859

860

861

862

863

864

865

866

867

868

869

870

871

872

873

874

875

876

877

878

879

880

881

882

883

884

885

886

887

888

889

890

891

892

893

894

895

896

897

898

899

900

901

902

903

904

905

906

907

908

909

910

911

912

913

914

915

916

917

918

919

920

921

922

923

924

925

926

927

928

929

930

931

932

933

934

935

936

937

938

939

940

941

942

943

944

945

946

947

948

949

950

951

952

953

954

CRITICAL REVIEWS IN BIOTECHNOLOGY 9

co-expression of heterologous RoMDH and RoFUM1failed to accumulate FA in the engineered strain. Itwas also observed that the fumarase deleted straincould produce FA via the oxidative route. Moreover, itwas observed that the introduction of the heterol-ogous reductive route of FA led to lower productionof FA but the introduction of a heterologous gene forpyruvate carboxylase (RoPYC) led to a higher FA titer(2.48 g/L). The maximum FA titer obtained, under opti-mal conditions of nitrogen limitation and the presenceof biotin (co-factor of PYC), was 5.6 g/L [5].

Another strategy investigated improved FA yield,investigated by the same group, was to subject the S.cerevisiae CEN.PK2-1C pdc1adh1fum1 strain to site-directed mutagenesis for improving the catalytic activ-ity of RoPYC as well as overexpression of the samestrain. The site-directed mutagenesis was able toimprove the catalytic activity of it by reducing theconformational instability afforded by the presence ofthe amino acid proline at 474 positions. Substitutionof the amino acid with asparagine led to improved FAproduction than the wild type strains. The best FAyield of 448.4mg/L was obtained following biotinsupplementation of 56 mg/L [71]. This yield is consid-erably lower than their previous study which yielded5.6 g/L fumarate [5]. Hence, a better understanding ofthe underlying mechanism of the metabolic pathwaysinfluencing fumarate production is essential for suc-cessful genetic modification strategies to improveits yield.

Torulopsis glabrata

Recently, multivitamin auxotrophic mutant strains ofT. glabrata have been investigated for fumaric acid pro-duction following genetic modifications. The auxo-trophic strains have been known to produce high titersof pyruvic acid and reported to have a higher glucoseand acid tolerance than S. cerevisiae [66]. Hence, isan ideal strain for use in the industrial production oforganic acids, mainly pyruvate, but has been subjectedto genetic modification to enable the production ofmalic acid as well as FA [66,72].

The genetic modification strategy adopted involvedidentification of the metabolic pathways that contrib-uted to the accumulation of FA in the cytosol fol-lowed by the introduction of enzymes involved in themetabolic pathways. To this effect, key enzymes affect-ing fumarate accumulation, involved in both carbonand nitrogen metabolism such as argininosuccinatelyase (ASL), adenylosuccinate lyase (ADSL), fumarylace-toacetase (FAA), and fumarase (FUM1), were identifiedusing in the silico metabolic model (iNX804). The

overexpression of the enzymes individually led to theaccumulation of FA in the modified strains comparedto the parent auxotrophic T. glabrata strain but ASLand ADSL were identified as the key enzymes leadingto higher FA accumulation. Optimization of overex-pression levels of enzymes ASL and ADSL gave a FAtiter of 5.62 g/L. This was further improved by intro-ducing a heterogeneous gene, SpMAE1, that encodesfor a C4-dicarboxylic acids transporter in S. pombe.The final titer obtained was 8.83 g/L [66].

Genetic modification of T. glabrata was performed inanother study which involved the introduction of heter-ogenous R. oryzae genes that code for the fumaraseenzyme (RoFUM), pyruvate carboxylase (RoPYC) andmalate dehydrogenase (RoMDH) as well as the C4-dicarboxylic transporter, SpMAE1, from S. pombe. Thebinding site B of RoFUM was subject to site-directedmutagenesis to enhance the catalytic efficiency of theenzyme. The resulting modified strain yielded a fumar-ate titer of 5.1 g/L. Further increases in FA yield wereachieved by subjecting the strain to additional modifi-cation. This additional modification involved deletionof the gene, ade12, that codes for adenylosuccinatesynthetase that is involved in purine metabolism. Theobtained strain resulted in a significantly higher FAyield of 9.2 g/L as well as increased acid tolerance [72].

Limitations of current genetic engineering strategies

The concentrations of the fumaric acid obtained usinggenetically engineered strains of E. coli, S. cerevisiae,and T. glabrata are significantly lower than thatreported for R. oryzae, this is summarized in Table 3.Despite the lower yields, genetic modification techni-ques are an important tool to improve to furtherimprove fumarate yield. These genetic and metabolicengineering strategies exhibit a special significance todecrease the cost of production as well as efficientutilization of biomass and industrial residues as feed-stocks for fumaric acid production. Moreover, thesemodifications allow for the alleviation of the problemsassociated with R. oryzae mediated FA bioproductiondiscussed in Limitations of R. oryzae based productionsection. Advanced genetic and metabolic engineeringstrategies that may be adopted for strain improve-ment are discussed in the following sections.

Advanced genetic and metabolic engineeringstrategies for fumaric acid production and itsderivatives

Advances in genetic and metabolic engineering havepermitted for modifications to inherent metabolic

955

956

957

958

959

960

961

962

963

964

965

966

967

968

969

970

971

972

973

974

975

976

977

978

979

980

981

982

983

984

985

986

987

988

989

990

991

992

993

994

995

996

997

998

999

1000

1001

1002

1003

1004

1005

1006

1007

1008

1009

1010

1011

1012

1013

1014

1015

1016

1017

1018

1019

1020

1021

1022

1023

1024

1025

1026

1027

1028

1029

1030

1031

1032

1033

1034

1035

1036

1037

1038

1039

1040

1041

1042

1043

1044

1045

1046

1047

1048

1049

1050

1051

1052

1053

1054

1055

1056

1057

1058

1059

1060

10 J. SEBASTIAN ET AL.

pathways to be carried out as well as the expressionof relevant heterologous pathways in desired micro-organisms. Few of such strategies that can be con-sidered for the production of fumaric acid as wellas its derivative esters are discussed in the follow-ing sections.

Genetic and metabolic engineering tools for thebioproduction of fumaric acid

Genetic transformation systems have allowed for theintroduction of foreign DNA in filamentous fungi, suchas R. oryzae, to produce strains with desired character-istics. Most of these transformation techniques, suchas plasmid vector-mediated transformation, producestrains with the transforming DNA remaining extrac-hromosomally leading to low mitotic stability [75].Moreover, alternate transformation techniques, such asprotoplast (spheroplast) fusion, are complicated aswell as inconsistent due to inherent variability in thecomposition of fungal cell walls of different strains[52]. To this effect, alternate transformation techniquesthat provide for stable chromosomal integration of theintroduced DNA are required.

Agrobacterium-mediated transformation (AMT)

For successful genetic transformation, an efficienttransformation technique is required to insert genesof interest. Agrobacterium tumefaciens-mediated trans-formation (ATMT) is a such technique and has beenused for the transformation of filamentous fungi. Themain advantage of this method is that it generatestransformants that have single genomic integration ofthe desired gene [76]. The soil bacterium A. tumefa-ciens has the ability to transfer a part of its DNA calledtransfer DNA (T-DNA) and this ability is exploited forheterologous protein expression in the host organism.For this purpose, the T-DNA is altered to contain the

gene of interest. After the introduction of the modi-fied T-DNA into the host by A. tumefaciens, the T-DNAintegrates into the chromosome of the host, therebyachieving the transformation of the host organism. Abrief summary of the technique and factors affectingAMT efficiency has been provided by Li et al. [52].

The filamentous fungus R. oryzae has been sub-jected to transformation using the AMT technique[75]. It was concluded that the technique was not suit-able for the expression of the gene of interest inR. oryzae due to the presence of a possible defensemechanism. However, more recent researches haveproved that heterologous gene expression is possible.The ectomycorrhizal fungus Laccaria laccata was suc-cessfully transformed through the ATM technique toexpress hygromycin B phosphotransferase (hph) [76].In another study, the filamentous fungus Trichodermareesei QM9414 was transformed using the AMT tech-nique which resulted in the formation of a hygromycinB-resistant fungi [77]. On the other hand, a decreasein resistance to hygromycin B was observed duringseveral cultivation cycles of the fungus Backusellalamprospora post successful integration of the heterol-ogous gene [78].

The AMT technique has been used for the geneticengineering of a diverse group of organisms but thereare no reports of using this technique for the trans-formation of fungi for organic acid production, espe-cially fumaric acid. The requirement of binary vectorsfor this method may be one of the probable reasons.These binary vector systems are tedious to prepare[79]. Also, there are numerous factors that play a keyrole in determining the transformation efficiencywhich needs to be optimized. This genetic modifica-tion and successful integration is a time-consumingprocess. Moreover, the presence of a possible defensemechanism against the presence of heterologousgenes may have contributed to the non-adoption ofthe AMT technique for the fungal transformation [75].

1061

1062

1063

1064

1065

1066

1067

1068

1069

1070

1071

1072

1073

1074

1075

1076

1077

1078

1079

1080

1081

1082

1083

1084

1085

1086

1087

1088

1089

1090

1091

1092

1093

1094

1095

1096

1097

1098

1099

1100

1101

1102

1103

1104

1105

1106

1107

1108

1109

1110

1111

1112

1113

1114

1115

1116

1117

1118

1119

1120

1121

1122

1123

1124

1125

1126

1127

1128

1129

1130

1131

1132

1133

1134

1135

1136

1137

1138

1139

1140

1141

1142

1143

1144

1145

1146

1147

1148

1149

1150

1151

1152

1153

1154

1155

1156

1157

1158

1159

1160

1161

1162

1163

1164

1165

1166

Table 3. Comparison of fumaric acid titer and yield of natural strains and genetically engineered strains.Strains Fumaric acid titer (g/L) Yield (g/g) Reference

Natural strainsR. arrhizus 103 0.79 [54]

98 0.82 [55]R. oryzae 56.2 0.54 [31]

56.5 0.94 [73]40.5 0.51 [74]

Engineered strainsE. coli CWF812 28.2 0.39 [26]E. coli EF02(pSCppc) 41.5 0.44 [27]R. oryzae ppc 25 0.78 [57]S. cerevisiae FMME 001 PYC2 þ RoMDH 3.18 0.05 [70]S. cerevisiae FMME 006 DFUM1 þ RoPYCþ RoMDHþ RoFUM1 5.64 0.11 [5]T. glabrata ASL(H) – ADSL(L) – SpMAE1 8.83 0.15 [66]T. glabrata (Dade12) – PMS – P160A 9.2 0.15 [72]

CRITICAL REVIEWS IN BIOTECHNOLOGY 11

The loss of introduced heterologous characteristicswith time further contributes to the unsuitability ofthe technique [78].

crispr cas-9

The advances in genetic engineering technologies havecreated a number of opportunities for basic biologicalresearch with filamentous fungi. Clustered regularlyinterspaced short palindromic repeats (CRISPR)-CRISPR-associated proteins (Cas) has recently emerged as apotent genetic engineering technique. It is a potentialcandidate to resolve the issue of low edit gene fre-quency observed in the case of filamentous fungi. Acomprehensive review of using the system for editingthe fungal genome is provided by Deng et al. [80] andShi et al. [81]. The efficient, simple, and fast techniqueof genetic modification allows for systemic researchon filamentous fungi. This technique is in its infancyand researchers are continuously learning about thepotential applications of this potent genome-edit-ing tool.

This genetic engineering technique has been mostlyused to perform gene deletion or creating a knockouttransformants. The versatility of the Cas9-multipleSgRNA allows for insertion, deletion, and replacementof target genes, thereby generating positive multiplegene modifications. From the fumaric acid productionpoint of view, the system can be used to efficiently pro-duce transformants with the deletion of genes thatencode enzymes, such as lactate dehydrogenase andalcohol dehydrogenase, that causes the production ofbyproducts. This enables the redirection of metabolicflux towards fumarate production [82].

Moreover, this system was used in the develop-ment of a genetic engineering toolbox for A. niger.The toolbox allows for the generation of strains car-rying heterologous expression cassettes at a definedgenetic locus. The proof of this concept was shownby the transformation of A. niger to produce aconiticacid. The transformed cells produced aconitate onlyin the presence of the inducer 10 mg/mL doxycycline[83]. Similar expression systems may be used for theproduction of fumaric acid whereas R. oryzae can betransformed to overexpress the enzymes involved inthe reductive pathway. Additionally, the enzymesinvolved in the glyoxylate pathway may be overex-pressed leading to the accumulation of succinateand malate which can then be converted to fumaricacid by succinate dehydrogenase and fumaraserespectively.

Gene knockout and silencing

Gene knockout and gene silencing have becomeimportant tools for genetic engineering. During geneknockout, the DNA is modified deliberately to removethe entire or part of the gene whereas gene silencing isa natural process where the expression is reduced. Theinterlinked nature of metabolic pathways, as repre-sented in Figure 3(a), can lead to the production ofmultiple byproducts in addition to the compound ofinterest. For example, during the synthesis of fumaratefrom the glucose, byproducts, such as ethanol, lacticacid, and glycerol, are also synthesized. This leads tothe loss or redirection of some of the glucose in theculture towards the synthesis of these byproducts. Theinactivation of the genes responsible for the synthesisof the enzymes, or the prevention of their synthesis,involved in the production of these byproducts canlead to the redirection of the metabolic pathwaytowards the synthesis of only the desired product, asdepicted in Figure 3(b).

For example, the inactivation of enzymes pyruvatedecarboxylase and lactate dehydrogenase can preventthe formation of the byproducts ethanol and lactate,respectively. This redirects the metabolic pathwaytowards fumarate synthesis. The inactivation of theenzymes involved can be achieved by the use oftechniques such as gene knockout or RNAi (RNA inter-ference) [84]. Gheinani et al. [85] were able to suc-cessfully silence the ldhA and ldhB genes of R. oryzaewith siRNA technology, the most commonly used RNAinterference (RNAi) tool and redirects metabolic activ-ity towards ethanol synthesis. Similar deactivationof the lactate dehydrogenase gene of an anaerobicgut fungus, Pecoramyces ruminantium strain C1Awas performed and reduced lactate production wasobserved [86].

Possible metabolic pathway modifications inR. oryzae

The modification of the metabolic pathway of E. coliby activation of the glyoxylate pathway has proved tobe successful in enabling the transformed strain toaccumulate FA [26,27]. Hence, this re-direction of themetabolic pathway in conjunction with the inherentreductive TCA cycle can be introduced in R. oryzae toenhance the production of FA. This possibility oftransformation has not been reported and the redir-ected metabolic pathway is represented in Figure 4.This redirection via the glyoxylate pathway hasthe potential to enhance FA accumulation in this fun-gal strain. It has been reported that the glyoxylate

1167

1168

1169

1170

1171

1172

1173

1174

1175

1176

1177

1178

1179

1180

1181

1182

1183

1184

1185

1186

1187

1188

1189

1190

1191

1192

1193

1194

1195

1196

1197

1198

1199

1200

1201

1202

1203

1204

1205

1206

1207

1208

1209

1210

1211

1212

1213

1214

1215

1216

1217

1218

1219

1220

1221

1222

1223

1224

1225

1226

1227

1228

1229

1230

1231

1232

1233

1234

1235

1236

1237

1238

1239

1240

1241

1242

1243

1244

1245

1246

1247

1248

1249

1250

1251

1252

1253

1254

1255

1256

1257

1258

1259

1260

1261

1262

1263

1264

1265

1266

1267

1268

1269

1270

1271

1272

12 J. SEBASTIAN ET AL.

pathway is required for fungal virulence [87]. Hence, acomprehensive understanding of the underlying riskof enhanced fungal virulence, if any, needs to bestudied extensively before any commercial applicationof the transformed strain is considered.

Additionally, metabolic modifications similar to thatdescribed by Chen et al. [66,72] can be investigatedwith R. oryzae. These modifications have been able to

increase FA titers in T. glabrata. Hence, the introduc-tion of metabolic enzymes involved in nitrogen andpurine metabolism are that carbon metabolism canbe considered to improve fumarate yield. Moreover,the introduction of an effective C4-dicarboxylicacid transporter, such as SpMAE1 and DcuB, that hasbeen shown to significantly improve fumarate produc-tion can be considered [66]. The acid tolerance of

1273

1274

1275

1276

1277

1278

1279

1280

1281

1282

1283

1284

1285

1286

1287

1288

1289

1290

1291

1292

1293

1294

1295

1296

1297

1298

1299

1300

1301

1302

1303

1304

1305

1306

1307

1308

1309

1310

1311

1312

1313

1314

1315

1316

1317

1318

1319

1320

1321

1322

1323

1324

1325

1326

1327

1328

1329

1330

1331

1332

1333

1334

1335

1336

1337

1338

1339

1340

1341

1342

1343

1344

1345

1346

1347

1348

1349

1350

1351

1352

1353

1354

1355

1356

1357

1358

1359

1360

1361

1362

1363

1364

1365

1366

1367

1368

1369

1370

1371

1372

1373

1374

1375

1376

1377

1378

Figure 3. (a) Metabolic pathways involved in the production of fumaric acid and byproducts and (b) reduction of byproduct for-mation by gene inactivation (represented in red) to obtain a high fumaric acid titer. gdp—Glyceraldehyde-3-phosphate dehydro-genase; ldh—lactate dehydrogenase, and pdc—pyruvate decarboxylase. For color version of the figure refer e-print.

CRITICAL REVIEWS IN BIOTECHNOLOGY 13

T. glabrata, as well as fumarate production, wasimproved by the deletion of the gene that codes foradenylosuccinate synthase (ade12). Similarly, themodification can be studied to improve the acid toler-ance of R. oryzae. This modification will have a dra-matic impact on the feasibility of commercial FAbioproduction strategies as the cost associated withthe use of the neutralizing agents as well as therecovery of FA from the fumarate complex formed bythe use of a neutralizing agent.

In situ biosynthesis of fumarate esters

The derivative of the fumaric acid, fumaric acid ester(FAEs), has recently been identified as an importantderivative due to its diverse medical applications.Hence, in situ bioconversion of fumaric acid to its esterforms provided an attractive alternative for the greensynthesis of the esters. The current methods for indus-trial synthesis of esters are nonselective, tedious, andconsume considerable energy. Most often these com-mercial processes rely on direct chemical esterification.

This high-temperature inorganic catalyst mediatedesterification reactions convert organic acids to estersin the presence of alcohol [88]. Due to the drawbacksof chemical processes, such as non-selectivity and highenergy consumption, biosynthesis has been proposedas a suitable alternative.

Lipases are the most used enzymes for the biosyn-thesis of a diverse group of pharmaceutical com-pounds. These enzymes have excellent stability evenin the presence of organic solvents, hence, their appli-cation in the production of diverse biosynthetic com-pounds. This application of enzymes in the synthesisof a diverse array of compounds has been reviewedextensively by Carvalho et al. [89]. R. oryzae is aknown producer of lipases and widely used for indus-trial applications [90,91]. Hence, it would be possibleto bioconvert FA produced to FAEs by inducing lipasesynthesis either via genetic modification or inductionof their synthesis. The alcohol needed for conversionof FA to FAEs may be provided externally. Ethanolproduction during FA production fermentation hasbeen reported [40]. The induction of this ethanol

1379

1380

1381

1382

1383

1384

1385

1386

1387

1388

1389

1390

1391

1392

1393

1394

1395

1396

1397

1398

1399

1400

1401

1402

1403

1404

1405

1406

1407

1408

1409

1410

1411

1412

1413

1414

1415

1416

1417

1418

1419

1420

1421

1422

1423

1424

1425

1426

1427

1428

1429

1430

1431

1432

1433

1434

1435

1436

1437

1438

1439

1440

1441

1442

1443

1444

1445

1446

1447

1448

1449

1450

1451

1452

1453

1454

1455

1456

1457

1458

1459

1460

1461

1462

1463

1464

1465

1466

1467

1468

1469

1470

1471

1472

1473

1474

1475

1476

1477

1478

1479

1480

1481

1482

1483

1484

Figure 4. Possible fumaric acid production via the introduction of the glyoxylate pathway (highlighted in blue) in conjunctionwith reductive TCA pathway (highlighted in blue) in R. oryzae. The various enzymes involved in the metabolic pathway havebeen represented in numbers. 1. Isocitrate lyase (EC 4.1.3.1); 2. malate synthase (EC 2.3.3.9); 3. pyruvate carboxylase (EC 6.4.1.1);4. malate dehydrogenase (EC 1.1.1.37); 5. fumarase (EC 4.2.1.2); 6. pyruvate dehydrogenase (EC 1.2.4.1); 7. citrate synthase (EC2.3.3.1); 8. aconitase (EC 4.2.1.3); 9. isocitrate dehydrogenase (EC 1.1.1.42); 10. 2-oxoglutarate dehydrogenase (EC 1.2.4.2); 11. suc-cinyl CoA synthase (EC 6.2.1.5); and 12. succinate dehydrogenase (complex III) (EC 1.3.99.1). For color version of the figure refere-print.

14 J. SEBASTIAN ET AL.

synthesis would not be recommended as the diversionof sugars to ethanol synthesis can lead to reducedFA production.

Conclusion

The depletion of the fossil fuels, their impact on theenvironment and the emphasis on green chemistryhas necessitated the production of microorganismmediated synthesis of platform chemicals. This shiftallows for the use of the renewable substrate as wellas industrial and domestic residues as feedstocksfor the manufacture of biochemicals. Due to the lowprice of commodity chemicals, such as fumaric acid,the main factors that determine economic viability isthe costs of substrate and downstream processing.These issues can be overcome by the use of biomassor organic residues as feedstocks for synthesis aswell as by employing efficient downstream processingstrategies. Hence, the main factor the determinesfeasibility is the product yield. Maximum product yieldis vital to lower manufacturing costs and to make bio-chemical production competitive against fossil fuel-based production.

Fumaric acid has been identified as one of the keybiobased platform chemicals that have diverse appli-cations. Studies to produce this dicarboxylic acidhave yielded high titers, especially by the fungus R.oryzae, but these high titers obtained are still lowerthan the theoretical expected yield. Moreover, thecomplications associated with the use of this fungalstrain have necessitated the development of an alter-native means for producing this acid. The advancesin metabolic engineering tools have allowed for thedevelopment of novel microorganisms capable toyield higher fumaric concentration compared to theparental strain. Some of these metabolic engineeringstrategies that have been investigated and can beadopted for improving FA yield has been discussed inthis review. The concentrations achieved are stilllower than those obtained by overproducing strainsof R. oryzae obtained from culture collections. Hence,more research particularly the optimization of themetabolic engineering strategies is essential to obtainhigher titers to make bioproduction feasible. Thisis particularly true of the redirection of metabolicpathways via the introduction of modifications thatenable in situ conversion of fumaric acid to its deriva-tives, such as fumarate esters and aspartic acid. Thiscan potentially improve the feasibility of the biopro-duction process.

Disclosure statement

No potential conflict of interest was reported by the authors.

Funding

The authors are sincerely thankful for the financial supportof the Minist�ere des Q12Relations Internationales du Qu�ebec(coop�eration Paran�a-Qu�ebec 2010–2012) and the NaturalSciences and Engineering Research Council of Canada(NSERC) Discovery Grant.

References

[1] Bozell JJ, Petersen GR. Technology development forthe production of biobased products from biorefinerycarbohydrates—the US Department of Energy’s “Top10” revisited. Green Chem. 2010;12. Q2

[2] Choi S, Song CW, Shin JH, et al. Biorefineries for theproduction of top building block chemicals and theirderivatives. Metab Eng. 2015;28:223–239.

[3] Roa Engel CA, Straathof AJJ, Zijlmans TW, et al.Fumaric acid production by fermentation. ApplMicrobiol Biotechnol. 2008;78:379–389.

[4] Rodr�ıguez-L�opez J, S�anchez AJ, G�omez DM, et al.Fermentative production of fumaric acid fromEucalyptus globulus wood hydrolyzates. J ChemTechnol Biotechnol. 2012;87:1036–1040.

[5] Xu G, Chen X, Liu L, et al. Fumaric acid production inSaccharomyces cerevisiae by simultaneous use of oxi-dative and reductive routes. Bioresour Technol. 2013;148:91–96.

[6] Martin-Dominguez V, Estevez J, Ojembarrena F, et al.Fumaric acid production: a biorefinery perspective.Fermentation. 2018;4. Q3

[7] Das RK, Brar SK, Verma M. Fumaric acid. Platformchemical biorefinery. Amsterdam, The Netherlands:Elsevier Inc.; 2016. Available from: http://linkinghub.elsevier.com/retrieve/pii/B9780128029800000080

[8] Das RK, Brar SK, Verma M. Recent advances in the bio-medical applications of fumaric acid and its esterderivatives: the multifaceted alternative therapeutics.Pharmacol Rep. 2016;68:404–414.

[9] Doscher CK, Kane JH, Cragwall GO, et al. Industrialapplications of fumaric acid. Ind Eng Chem. 1941;33:315–319.

[10] Reich K, Thaci D, Mrowietz U, et al. Efficacy and safetyof fumaric acid esters in the long-term treatment ofpsoriasis-“A retrospective study (FUTURE). J DtschDermatol Ges. 2009;7:603–610.

[11] Ermis U, Weis J, Schulz JB. PML in a patient treatedwith fumaric acid. N Engl J Med. 2013;368:1657–1658.

[12] Jansen J, Boerakker MJ, Heuts J, et al. Rapid photo-crosslinking of fumaric acid monoethyl ester-function-alized poly(trimethylene carbonate) oligomers fordrug delivery applications. J Controlled Release. 2010;147:54–61.

[13] Jansen J, Tibbe MP, Mihov G, et al. Photo-crosslinkednetworks prepared from fumaric acid monoethylester-functionalized poly(d,l-lactic acid) oligomers and

1485

1486

1487

1488

1489

1490

1491

1492

1493

1494

1495

1496

1497

1498

1499

1500

1501

1502

1503

1504

1505

1506

1507

1508

1509

1510

1511

1512

1513

1514

1515

1516

1517

1518

1519

1520

1521

1522

1523

1524

1525

1526

1527

1528

1529

1530

1531

1532

1533

1534

1535

1536

1537

1538

1539

1540

1541

1542

1543

1544

1545

1546

1547

1548

1549

1550

1551

1552

1553

1554

1555

1556

1557

1558

1559

1560

1561

1562

1563

1564

1565

1566

1567

1568

1569

1570

1571

1572

1573

1574

1575

1576

1577

1578

1579

1580

1581

1582

1583

1584

1585

1586

1587

1588

1589

1590

CRITICAL REVIEWS IN BIOTECHNOLOGY 15

Author Query
AQ3: Please provide the missing page range for the "[6]" references list entry.
Author Query
AQ2: Please provide the missing page range for the "[1]" references list entry.
Author Query
AQ12: The funding information provided has been checked against the Open Funder Registry and we failed to find a match. Please check and resupply the funding details if necessary.

N-vinyl-2-pyrrolidone for the controlled and sustainedrelease of proteins. Acta Biomater. 2012;8:3652–3659.

[14] Felthouse TR, Burnett JC, Horrell B, et al. Maleicanhydride, maleic acid, and fumaric acid. 4th ed. In:Kirk-Othmer Encyclopedia of Chemical Technology.Vol.19; 2001; p. 1–58.Q4

[15] Papadaki A, Androutsopoulos N, Patsalou M, et al.Biotechnological production of fumaric acid: theeffect of morphology of Rhizopus arrhizus NRRL 2582.Fermentation. 2017;3.Q5

[16] GrandViewResearch.com [Internet]. San Francisco[cited 2018 May 10]. Available from: https://www.grandviewresearch.com/industry-analysis/fumaric-acid-market.Q6

[17] IHSMarkit.com [Internet]. London [cited 2018 May10]. Available from: https://ihsmarkit.com/products/fumaric-acid-chemical-economics-handbook.html.Q7

[18] Goldberg I, Rokem JS, Pines O. Organic acids: oldmetabolites, new themes. J Chem Technol Biotechnol.2006;81:1601–1611.

[19] Xu Q, Li S, Huang H, et al. Key technologies for theindustrial production of fumaric acid by fermentation.Biotechnol Adv. 2012;30:1685–1696.

[20] Lee JW, Kim HU, Choi S, et al. Microbial productionof building block chemicals and polymers. Curr OpinBiotechnol. 2011;22:758–767.

[21] Sauer M, Porro D, Mattanovich D, et al. Microbial pro-duction of organic acids: expanding the markets.Trends Biotechnol. 2008;26:100–108.

[22] Roa Engel CA, van Gulik WM, Marang L, et al.Development of a low pH fermentation strategy forfumaric acid production by Rhizopus oryzae. EnzymeMicrob Technol. 2011;48:39–47.

[23] Zhang K, Zhang L, Yang S-T. Fumaric acid recoveryand purification from fermentation broth by activatedcarbon adsorption followed with desorption by acet-one. Ind Eng Chem Res. 2014;53:12802–12808.

[24] Figueira D, Cavalheiro J, Ferreira B. Purification ofpolymer-grade fumaric acid from fermented spent sul-fite liquor. Fermentation. 2017;3.Q8

[25] Yu S, Huang D, Wen J, et al. Metabolic profiling of aRhizopus oryzae fumaric acid production mutant gen-erated by femtosecond laser irradiation. BioresourTechnol. 2012;114:610–615.

[26] Song CW, Kim DI, Choi S, et al. Metabolic engineeringof Escherichia coli for the production of fumaric acid.Biotechnol Bioeng. 2013;110:2025–2034.

[27] Li N, Zhang B, Wang Z, et al. Engineering Escherichiacoli for fumaric acid production from glycerol.Bioresour Technol. 2014;174:81–87.

[28] Skory CD, Ibrahim AS. Native and modified lactatedehydrogenase expression in a fumaric acid produc-ing isolate Rhizopus oryzae 99-880. Curr Genet. 2007;52:23–33.

[29] Wakai S, Arazoe T, Ogino C, et al. Future insights infungal metabolic engineering. Bioresour Technol.2017;245:1314–1326.

[30] Abe A, Oda Y, Asano K, et al. Rhizopus delemar is theproper name for Rhizopus oryzae fumaric-malic acidproducers. Mycologia. 2007;99:714–722.

[31] Fu Y-Q, Li S, Chen Y, et al. Enhancement of fumaricacid production by Rhizopus oryzae using a two-stage

dissolved oxygen control strategy. Appl BiochemBiotechnol. 2010;162:1031–1038.

[32] Huang L, Wei P, Zang R, et al. High-throughputscreening of high-yield colonies of Rhizopus oryzaefor enhanced production of fumaric acid. AnnMicrobiol. 2010;60:287–292.

[33] Kang SW, Lee H, Kim D, et al. Strain development andmedium optimization for fumaric acid production.Biotechnol Bioproc E. 2010;15:761–769.

[34] Wen S, Liu L, Nie KL, et al. Enhanced fumaric acid pro-duction by fermentation of xylose using a modifiedstrain of Rhizopus arrhizus. BioResources 2013;8. Q9

[35] Liu H, Wang W, Deng L, et al. High production offumaric acid from xylose by newly selected strainRhizopus arrhizus RH 7-13-9#. Bioresour Technol. 2015;186:348–350.

[36] Liu H, Hu H, Jin Y, et al. Co-fermentation of a mixtureof glucose and xylose to fumaric acid by Rhizopusarrhizus RH 7-13-9#. Bioresour Technol. 2017;233:30–33.

[37] Xu Q, Li S, Fu Y, et al. Two-stage utilization of cornstraw by Rhizopus oryzae for fumaric acid production.Bioresour Technol. 2010;101:6262–6264.

[38] Deng Y, Li S, Xu Q, et al. Production of fumaric acidby simultaneous saccharification and fermentationof starchy materials with 2-deoxyglucose-resistantmutant strains of Rhizopus oryzae. Bioresour Technol.2012;107:363–367.

[39] Das RK, Brar SK. Enhanced Fumaric acid productionfrom brewery wastewater and insight into themorphology of Rhizopus oryzae 1526. Appl BiochemBiotechnol. 2014;172:2974–2988.

[40] Das RK, Brar SK, Verma M. A fermentative approachtowards optimizing directed biosynthesis of fumaricacid by Rhizopus oryzae 1526 utilizing apple industrywaste biomass. Fungal Biol. 2015;119:1279–1290.

[41] Liao W, Liu Y, Frear C, et al. Co-production of fumaricacid and chitin from a nitrogen-rich lignocellulosicmaterial – dairy manure – using a pelletized filament-ous fungus Rhizopus oryzae ATCC 20344. BioresourTechnol. 2008;99:5859–5866.

[42] Zhou Y, Nie K, Zhang X, et al. Production of fumaricacid from biodiesel-derived crude glycerol by Rhizopusarrhizus. Bioresour Technol. 2014;163:48–53.

[43] Gu C, Zhou Y, Liu L, et al. Production of fumaric acidby immobilized Rhizopus arrhizus on net. BioresourTechnol. 2013;131:303–307.

[44] Naude A, Nicol W. Fumaric acid fermentation withimmobilised Rhizopus oryzae: quantifying time-depend-ent variations in catabolic flux. Process Biochem. 2017;56:8–20.

[45] Liu H, Zhao S, Jin Y, et al. Production of fumaric acidby immobilized Rhizopus arrhizus RH 7-13-9# on loo-fah fiber in a stirred-tank reactor. Bioresour Technol.2017;244:929–933.

[46] Das RK, Brar SK, Verma M. Enhanced fumaric acidproduction from brewery wastewater by immobiliza-tion technique. J Chem Technol Biotechnol. 2015;90:1473–1479.

[47] Cao N, Du J, Gong CS, et al. Simultaneous productionand recovery of fumaric acid from immobilizedRhizopus oryzae with a rotary biofilm contactor and

1591

1592

1593

1594

1595

1596

1597

1598

1599

1600

1601

1602

1603

1604

1605

1606

1607

1608

1609

1610

1611

1612

1613

1614

1615

1616

1617

1618

1619

1620

1621

1622

1623

1624

1625

1626

1627

1628

1629

1630

1631

1632

1633

1634

1635

1636

1637

1638

1639

1640

1641

1642

1643

1644

1645

1646

1647

1648

1649

1650

1651

1652

1653

1654

1655

1656

1657

1658

1659

1660

1661

1662

1663

1664

1665

1666

1667

1668

1669

1670

1671

1672

1673

1674

1675

1676

1677

1678

1679

1680

1681

1682

1683

1684

1685

1686

1687

1688

1689

1690

1691

1692

1693

1694

1695

1696

16 J. SEBASTIAN ET AL.

Author Query
AQ9: Please provide the missing page range for the "[34]" references list entry.
Author Query
AQ8: Please provide the missing page range for the "[24]" references list entry.
Author Query
AQ7: Please provide the missing article title for the "[17]" references list entry.
Author Query
AQ6: Please provide the missing article title for the "[16]" references list entry.
Author Query
AQ5: Please provide the missing page range for the "[6]" references list entry.
Author Query
AQ4: Kindly provide the missing editor names and publisher name and location for the “[14} references list entry.

an adsorption column. Appl Environ Microbiol. 1996;62:2926–2931.

[48] Gangl IC, Weigand WA, Keller FA. Economic compari-son of calcium fumarate and sodium fumarate pro-duction byRhizopus arrhizus. Appl Biochem Biotechnol.1990;24-25:663–677.

[49] Zhou Z, Du G, Hua Z, et al. Optimization of fumaricacid production by Rhizopus delemar based on themorphology formation. Bioresour Technol. 2011;102:9345–9349.

[50] Ding Y, Li S, Dou C, et al. Production of fumaric acidby Rhizopus oryzae: role of carbon–nitrogen ratio.Appl Biochem Biotechnol. 2011;164:1461–1467.

[51] Meussen BJ, de Graaff LH, Sanders JPM, et al.Metabolic engineering of Rhizopus oryzae for theproduction of platform chemicals. Appl MicrobiolBiotechnol. 2012;94:875–886.

[52] Li D, Tang Y, Lin J, et al. Methods for genetic trans-formation of filamentous fungi. Microb Cell Fact.2017;16.Q10

[53] Dolatabadi S, de Hoog GS, Meis JF, et al. Speciesboundaries and nomenclature of Rhizopus arrhizus(-syn. R. oryzae). Mycoses 2014;57:108–127.

[54] Rhodes RA, Lagoda AA, Misenheimer TJ, et al.Production of fumaric acid in 20-liter fermentors. ApplMicrobiol. 1962; 10:9–15.

[55] Kenealy W, Zaady E, Preez JC d, et al. Biochemicalaspects of fumaric acid accumulation by Rhizopusarrhizus. Appl Environ Microbiol. 1986; 52:128–133.

[56] Fu Y-Q, Xu Q, Li S, et al. Strain improvement ofRhizopus oryzae for over-production of fumaric acidby reducing ethanol synthesis pathway. Korean JChem Eng. 2010;27:183–186.

[57] Zhang B, Skory CD, Yang S-T. Metabolic engineeringof Rhizopus oryzae: effects of overexpressing pyc andpepc genes on fumaric acid biosynthesis from glu-cose. Metab Eng. 2012;14:512–520.

[58] Zhang B, Yang S-T. Metabolic engineering of Rhizopusoryzae: effects of overexpressing fumR gene on cellgrowth and fumaric acid biosynthesis from glucose.Process Biochem. 2012;47:2159–2165.

[59] Peleg Y, Battat E, Scrutton M, et al. IsoenzymeQ11 patternand subcellular localisation of enzymes involved infumaric acid accumulation by Rhizopus oryzae. ApplMicrobiol Biotechnol. 1989;32.

[60] Brown SH, Bashkirova L, Berka R, et al. Metabolicengineering of Aspergillus oryzae NRRL 3488 forincreased production of l-malic acid. Appl MicrobiolBiotechnol. 2013;97:8903–8912.

[61] Lee SY, Lee D-Y, Kim TY. Systems biotechnologyfor strain improvement. Trends Biotechnol. 2005;23:349–358.

[62] Toya Y, Shimizu H. Flux analysis and metabolomics forsystematic metabolic engineering of microorganisms.Biotechnol Adv. 2013;31:818–826.

[63] Longacre A, Reimers JM, Gannon JE, et al. Flux ana-lysis of glucose metabolism in Rhizopus oryzaefor thepurpose of increasing lactate yields. Fungal GenetBiol. 1997;21:30–39.

[64] Bai D-M, Zhao X-M, Li X-G, et al. Strain improvementof Rhizopus oryzae for over-production of l(þ)-lactic

acid and metabolic flux analysis of mutants. BiochemEng J. 2004;18:41–48.

[65] Liu Y, Xu Q, Lv C, et al. Study of metabolic profile ofRhizopus oryzae to enhance fumaric acid productionunder low pH condition. Appl Biochem Biotechnol.2015;177:1508–1519.

[66] Chen X, Wu J, Song W, et al. Fumaric acid productionby Torulopsis glabrata: engineering the urea cycleand the purine nucleotide cycle. Biotechnol Bioeng.2015;112:156–167.

[67] Raab AM, Lang C. Oxidative versus reductive succinicacid production in the yeast Saccharomyces cerevisiae.Bioeng Bugs 2011;2:120–123.

[68] Zhang T, Wang Z, Deng L, et al. Pull-in urea cyclefor the production of fumaric acid in Escherichia coli.Appl Microbiol Biotechnol. 2015;99:5033–5044.

[69] Zhang T, Song R, Wang M, et al. Regulating C4-dicar-boxylate transporters for improving fumaric acid pro-duction. RSC Adv. 2017;7:2897–2904.

[70] Xu G, Liu L, Chen J. Reconstruction of cytosolicfumaric acid biosynthetic pathways in Saccharomycescerevisiae. Microb. Cell Fact. 2012;11. Q13

[71] Xu G, Wu M, Jiang L. Site-saturation engineering ofproline 474 in pyruvate carboxylase from Rhizopusoryzae to elevate fumaric acid production in engi-neered Saccharomyces cerevisiae cells. Biochem Eng J.2017;117:36–42.

[72] Chen X, Song W, et al. Fumarate production byTorulopsis glabrata: engineering heterologous fumar-ase expression and improving acid tolerance. PLoSOne. 2016;11. Q14

[73] Wang G, Huang D, Qi H, et al. Rational medium opti-mization based on comparative metabolic profilinganalysis to improve fumaric acid production. BioresourTechnol. 2013;137:1–8.

[74] Gu S, Xu Q, Huang H, et al. Alternative respiration andfumaric acid production of Rhizopus oryzae. ApplMicrobiol Biotechnol. 2014;98:5145–5152.

[75] Michielse CB, Salim K, Ragas P, et al. Development ofa system for integrative and stable transformation ofthe zygomycete Rhizopus oryzae by Agrobacterium-mediated DNA transfer. Mol Genet Genomics. 2004;271:499–510.

[76] Zubieta MP, da Silva Coelho I, de Queiroz MV, et al.Agrobacterium tumefaciens-mediated genetic trans-formation of the ectomycorrhizal fungus Laccaria lac-cata. Ann Microbiol. 2014;64:1875–1878.

[77] Zhong YH, Wang XL, Wang TH, et al. Agrobacterium-mediated transformation (AMT) of Trichoderma reeseias an efficient tool for random insertional mutagen-esis. Appl Microbiol Biotechnol. 2007;73:1348–1354.

[78] Nyilasi I, Papp T, Csernetics �A, et al. Agrobacteriumtumefaciens-mediated transformation of the zygomy-cete fungus Backusella lamprospora. J Basic Microbiol.2008;48:59–64.

[79] Lee LY, Gelvin SB. T-DNA binary vectors and systems.Plant Physiol. 2007;146:325–332.

[80] Deng H, Gao R, Liao X, et al. CRISPR system in fila-mentous fungi: current achievements and futuredirections. Gene. 2017;627:212–221.

1697

1698

1699

1700

1701

1702

1703

1704

1705

1706

1707

1708

1709

1710

1711

1712

1713

1714

1715

1716

1717

1718

1719

1720

1721

1722

1723

1724

1725

1726

1727

1728

1729

1730

1731

1732

1733

1734

1735

1736

1737

1738

1739

1740

1741

1742

1743

1744

1745

1746

1747

1748

1749

1750

1751

1752

1753

1754

1755

1756

1757

1758

1759

1760

1761

1762

1763

1764

1765

1766

1767

1768

1769

1770

1771

1772

1773

1774

1775

1776

1777

1778

1779

1780

1781

1782

1783

1784

1785

1786

1787

1788

1789

1790

1791

1792

1793

1794

1795

1796

1797

1798

1799

1800

1801

1802

CRITICAL REVIEWS IN BIOTECHNOLOGY 17

Author Query
AQ14: Please provide the missing page range for the "[72]" references list entry.
Author Query
AQ13: Please provide the missing page range for the "[70]" references list entry.
Author Query
AQ11: Please provide the missing page range for the "[59]" references list entry.
Author Query
AQ10: Please provide the missing page range for the "[52]" references list entry.

[81] Shi T-Q, Liu G-N, Ji R-Y, et al. CRISPR/Cas9-based gen-ome editing of the filamentous fungi: the state of theart. Appl Microbiol Biotechnol. 2017;101:7435–7443.

[82] Nødvig CS, Hoof JB, Kogle ME, et al. Efficient oligonucleotide mediated CRISPR-Cas9 gene editing inAspergilli. Fungal Genet Biol. 2018;115:78–89.

[83] Sarkari P, Marx H, Blumhoff ML, et al. An efficient toolfor metabolic pathway construction and gene integra-tion for Aspergillus niger. Bioresour Technol. 2017;245:1327–1333.

[84] Whitehead KA, Langer R, Anderson DG. Knockingdown barriers: advances in siRNA delivery. Nat RevDrug Discov. 2009;8:129–138.

[85] Gheinani AH, Jahromi NH, Feuk-Lagerstedt E, et al. RNAsilencing of lactate dehydrogenase gene in Rhizopusoryzae. J RNAi Gene Silencing. 2011;7:443–448.

[86] Calkins SS, Elledge NC, Mueller KE, et al. Developmentof an RNA interference (RNAi) gene knockdown

protocol in the anaerobic gut fungus Pecoramycesruminantium strain C1A. PeerJ. 2018;6. Q15

[87] Lorenz MC, Fink GR. The glyoxylate cycle is requiredfor fungal virulence. Nature. 2001;412:83–86.

[88] Pacheco BS, Nunes CFP, Rockembach CT, et al. Eco-friendly synthesis of esters under ultrasound withp-toluenesulfonic acid as catalyst. Green Chem LettRev. 2014;7:265–270.

[89] Carvalho A, Fonseca T, Mattos M, et al. Recent advan-ces in lipase-mediated preparation of pharmaceuticalsand their intermediates. Int J Mol Sci. 2015;16:29682–29716.

[90] Salah RB, Mosbah H, Fendri A, et al. Biochemical andmolecular characterization of a lipase produced byRhizopus oryzae. FEMS Lett. 2006;260:241–248.

[91] Witt SN, Satomura A, Kuroda K, et al. Generation of afunctionally distinct Rhizopus oryzae lipase throughprotein folding memory. PLos One. 2015;10. Q16

1803

1804

1805

1806

1807

1808

1809

1810

1811

1812

1813

1814

1815

1816

1817

1818

1819

1820

1821

1822

1823

1824

1825

1826

1827

1828

1829

1830

1831

1832

1833

1834

1835

1836

1837

1838

1839

1840

1841

1842

1843

1844

1845

1846

1847

1848

1849

1850

1851

1852

1853

1854

1855

1856

1857

1858

1859

1860

1861

1862

1863

1864

1865

1866

1867

1868

1869

1870

1871

1872

1873

1874

1875

1876

1877

1878

1879

1880

1881

1882

1883

1884

1885

1886

1887

1888

1889

1890

1891

1892

1893

1894

1895

1896

1897

1898

1899

1900

1901

1902

1903

1904

1905

1906

1907

1908

18 J. SEBASTIAN ET AL.

Author Query
AQ16: Please provide the missing page range for the "[91]" references list entry.
Author Query
AQ15: Please provide the missing page range for the "[86]" references list entry.