Metabolic Engineering of Clostridium tyrobutyricum for Production of Biofuels and Bio-based Chemicals

Dissertation

Presented in Partial Fulfillment of the Requirement for the Degree Doctor of Philosophy

in the Graduate School of The Ohio State University

By

Yali Zhang, M.S.

Ohio State Biochemistry Program

The Ohio State University

2009

Dissertation Committee

Professor Shang-Tian Yang, Adviser Approved by

Professor Hua Wang ______

Professor Peng George Wang Adviser

Professor Jeffrey Chalmers

Copyright by

Yali Zhang

2009 ABSTRACT

The depletion of petroleum resource is aggravated by the increasing demand for fossil fuels. Moreover, the burning process of fossil fuels causes global warming and pollution. These concerns promote the intensive investigation on the development of economical processes to produce fuels and chemicals based on renewable resources. In this respect, plant biomass is the current sustainable source for the production of bio- based chemicals and biofuels.

C. tyrobutyricum is a rod-shape, gram-positive bacterium that produces butyrate, acetate, and hydrogen from various saccharides, including glucose and xylose, under strict anaerobic conditions. Phosphotransbutyrylase (PTB) is a key in the synthesis pathway. However, its role in affecting metabolic flux distribution and production of various metabolites is not well understood. In this work, a mutant with inactivated ptb gene, encoding phosphotransbutyrylase, was created by integrational mutagenesis through homologous recombination. Compared to the wild-type, the activities of phosphotransbutyrylase (PTB) and butyrate (BK) in the mutant decreased 76% and 42%, respectively; meanwhile, phosphotransacetylase (PTA) and

ii (AK) increased 7% and 29%, respectively. In addition, the mutant displayed a higher tolerance to butyric acid.

The fermentation conditions were optimized for the ptb mutant fermentation. The pH value was set at 6 and the initial concentration was 50 g/l by using one- factor-a-time screening. The impact of ptb-disruption on the fermentation profile was studied using glucose and xylose as substrate. Compared to the wild type, the b/a ratio decreased by 34.7% for glucose and 28.1% for xylose, this could be attributed to the inactivation of ptb. The productivity of hydrogen increased 38% for glucose and 46% for xylose, which could be partially attributed to a higher specific growth rate (33% increase for glucose and 40% increase for xylose) in addition to the lower b/a ratio. The butyric acid fermentation and hydrogen production were further improved by immobilizing the cells in the fibrous-bed bioreactor to facilitate cell adaptation to attain a higher final concentration.

The genome of C. tyrobutyricum was sequenced by using the 454 technology.

Several genes encoding involved in the solvent formation pathway were annotated. Interestingly, the only missing gene for a complete butanol production route is for butyrylaldehyde dehydrogenase. Moreover, the mutants of C. tyrobutyricum showed higher tolerance to butanol than the butanol producing C. acetobutylicum. Plasmids carrying aldehyde/ dehydrogenase gene aad or adhE2 from C. acetobutylicum

iii were constructed and introduced into C. tyrobutyricum. The engineered mutants showed significant butanol production in batch fermentation with glucose as the substrate. This work demonstrated that it is feasible to use C. tyrobutyricum to produce butanol.

iv

Dedicated to my parents and boyfriend Bingxu Song

v ACKNOWLEDGEMENTS

First and foremost, I would like to thank my advisor, Dr. Shang-Tian Yang, for

his guidance, encouragement, patience, and support during my graduate study. I am

extremely grateful for the flexibility he gave me during my research. I learned not only in

scientific knowledge, but also in his great personality. I will undoubtedly benefit from

these throughout my lifetime.

I also wish to express my sincere thanks to Dr. Hua Wang, Dr. Peng George

Wang and Dr. Jeffrey Chalmers for their constructive suggestions and advice offered on a

variety of topics.

I would like to acknowledge members of our group, Dr. An Zhang, Ms. Baohua

Zhang, and Ms. Wei-Lun Chang, for their helpful discussion and comments. Special

thanks go to Dr. Mingrui Yu and Ms. Milky Agarwal for their help in the work reported

in Chapters 5. The technical advice and assistance provided by Dr. Xiaoguang Liu at the initial stage of my research is gratefully acknowledged. Thanks are extended to Ms.

Jianxin Sun and Ms. Yanyan Zhang for their friendship and encouragement.

Financial supports from the Ohio Department of Development Third Frontier

Advanced Energy Program, the Consortium for Plant Biotechnology Research, Inc., and

EnerGenetics International during the course of this research are acknowledged.

vi At last, but most importantly, I wish to thank my parents, my aunt Dangsheng

Zhang and my boyfriend Bingxu Song for their unconditional love and confidence in my ability. All the other family members and all of my friends are acknowledged for their encouragement and support.

vii VITA

February 12, 1976 ...... Born-Datong, China September 1993-July 1997 ...... B.S. Chemistry, Nankai University, Tianjin, China September 1997-June 2000 ...... Research Associate, Shanghai Institute of Pharmaceutical Industry, Shanghai, China

January 2003-June 2004 ...... M.S. Biotechnology, The Ohio State University September 2004-August 2005 ...... Fellowship, Ohio State Biochemistry Program The Ohio State University

September 2005-present ...... Graduate Research Associate, Chemical & Biomolecular Engineering, The Ohio State University

PUBLICATIONS

1. Shang-Tian Yang, Xiaoguang Liu, and Yali Zhang, Metabolic engineering: applications, methods, and challenges, in S.T. Yang (ed.), Bioprocessing for Value-Added Products from Renewable Resources, Elsevier (2006), pp. 73-118

FIELDS OF STUDY

Major Field: Biochemistry Specialty: Biochemical Engineering

viii Table of Contents

Abstract ...... ii Dedication ...... v Acknowledgements ...... vi Vita ...... viii List of Figures ...... xv List of Tables ...... xviii Chapters

1. Introduction ...... 1 1.1 Objectives ...... 3 1.2 References ...... 7 2. Literature Review ...... 11 2.1 Bio-based chemicals production ...... 11 2.1.1 Biomass ...... 11 2.1.2 Bio-based chemicals ...... 13 2.1.2.1 Hydrogen...... 13 2.1.2.2 Butyric acid ...... 17 2.1.2.3 Butanol ...... 20 2.2 Metabolic engineering and strain development ...... 23 2.2.1 Clostridium ...... 23 2.2.2 Metabolic engineering ...... 25 2.2.3 Key enzymes related to the study ...... 27 2.2.3.1 PTB and BK ...... 28

ix 2.2.3.2 PTA and ACK ...... 29 2.2.3.3 Hydrogenase ...... 29 2.2.3.4 Aldehyde/alcohol dehydrogenase ...... 30 2.3 Process development ...... 31 2.3.1 Medium composition ...... 31 2.3.1.1 Substrate selection and initial concentration ...... 31 2.3.1.2 pH effect ...... 34 2.3.2 Fermentation mode ...... 34 2.3.2.1 Immobilized cell ...... 34 2.3.2.2 Continuous fermentation ...... 36 2.3.2.3 Sparging and operational controls ...... 37 2.4 Genome sequence and microarray ...... 37 2.4.1 Genome sequence ...... 37 2.4.2 Microarray...... 38 2.5 References ...... 40 3. Construction and characterization of ptb- mutant of Clostridium tyrobutyricum ...... 59 Summary ...... 59 3.1 Introduction ...... 60 3.2 Materials and methods ...... 62 3.2.1 Cultures and media ...... 62 3.2.2 Cloning of ptb deleted mutant ...... 62 3.2.2.1 PCR amplification for ptb fragment ...... 63 3.2.2.2 Nested-PCR for the confirmation of C. tyrobutyricum chromosome...... 63 3.2.2.3 Construction of non-replicative plasmid pPtbEm ...... 64

x 3.2.2.4 Transformation ...... 65 3.2.2.5 PCR characterization of ptb-disrupted mutant ...... 65 3.2.3 Enzyme activity assays ...... 66 3.2.4 Butyrate tolerance ...... 67 3.2.5 Fermentation kinetics ...... 68 3.2.6 Analytical methods ...... 68 3.3 Results ...... 68 3.3.1 Cloning of ptb deleted mutant ...... 68 3.3.2 Effects of ptb-disruption on acid forming enzyme activities ...... 70 3.3.3 Effect on fermentation profile ...... 71 3.3.4 butyrate tolerance ...... 71 3.4 Discussion ...... 72 3.5 Conclusion ...... 76 3.6 References ...... 77 4. Kinetics of butyric acid fermentation by Clostridium tyrobutyricum ...... 91 Summary ...... 91 4.1 Introduction ...... 92 4.2 Materials and methods ...... 94 4.2.1 Cultures and media ...... 94 4.2.2 Design of experiments for medium composition ...... 95 4.2.3 Effect of initial glucose concentration ...... 96 4.2.4 Effect of pH ...... 96 4.2.5 Free cell fermentations ...... 97 4.2.6 Immobilized cell fermentation in FBB ...... 97 4.2.7 Analytical methods ...... 98 4.3 Results ...... 99

xi 4.3.1 Simplified medium composition ...... 99 4.3.2 Substrates selection ...... 102 4.3.3 Optimized fermentation conditions ...... 102 4.3.4 Effects of ptb knock-out on the fermentation kinetics ...... 103 4.3.5 Free cell fed-batch fermentation kinetics ...... 104 4.3.6 FBB fed-batch fermentation kinetics ...... 105 4.4 Discussion ...... 105 4.5 Conclusion ...... 108 4.6 References ...... 110 5. Genome sequencing and metabolic engineering of C. tyrobutyricum for butanol production ...... 127 Summary ...... 127 5.1 Introduction ...... 127 5.2 Materials and methods ...... 130 5.2.1 Strains, plasmids and media ...... 130 5.2.2 Genome DNA extraction ...... 131 5.2.3 454 sequencing and metabolic pathway reconstruction...... 131 5.2.4 pCAAD and pSOS-adhE2 construction ...... 132 5.2.5 Restriction endonuclease system identification ...... 133 5.2.6 Transformation ...... 134 5.2.7 Butanol and butyrate tolerance ...... 134 5.2.8 Fermentation kinetics ...... 135 5.2.9 Analytical methods ...... 136 5.3 Results ...... 137 5.3.1 Genome sequencing ...... 137 5.3.2 Restriction endonuclease system identification ...... 138

xii 5.3.3 Transformation ...... 138 5.3.4 Butanol tolerance ...... 139 5.3.5 Fermentation kinetics ...... 140 5.4 Discussion ...... 141 5.5 Conclusion ...... 145 5.6 References ...... 146 6. Conclusions and recommendations...... 162 6.1 Conclusions ...... 162 6.1.1 Metabolic engineering ...... 162 6.1.2 Fermentation kinetics ...... 163 6.1.3 Genome sequence ...... 164 6.1.4 Butanol production by C. tyrobutyricum ...... 165 6.2 Recommendations ...... 165 6.2.1 Butanol production...... 165 6.2.2 CoA ...... 167 6.2.3 Functional genomics and proteomics ...... 168 Bibliography ...... 172 Appendices ...... 196 Appendix A Microarray fabrication ...... 196 A.1 Genomic library preparation ...... 196 A.2 Microarray slide preparation and quality control ...... 197 A.3 RNA extraction and cDNA preparation, dye labeling and competitive hybridization ...... 198 A.4 Scanning and data analysis ...... 199 Appendix B Summary of CoA transferase and Genes involved in glycolysis .....205 Appendix C Chromatography methods and standard spectrum ...... 210

xiii C.1 High performance liquid chromatography ...... 210 C.2 Gas chromatography for gas production ...... 210 C.3 Gas chromatography for alcohol analysis ...... 211

xiv List of Figures

Figure Page

1.1 The metabolic pathway in Clostridium tyrobutyricum ATCC 25755 ...... 9 1.2 Research objectives and scope of this study ...... 10 2.1 The metabolic pathway in Clostridium acetobutylicum ATCC824 ...... 57 2.2 The metabolic pathway in Clostridium tyrobutyricum ATCC25755 ...... 58 3.1 The metabolic pathway showing butyric acid, acetic acid, and hydrogen production from glucose in Clostridium tyrobutyricum ...... 82 3.2 Alignment of amino acid sequences of ptb of several Clostridium species showing highly conserved regions ...... 83 3.3 The nested-PCR to conform the chromosome from C. tyrobutyricum ...... 84 3.4 Construction of the non-replicative integrational plamid pPtbEm with one 0.64 kb ptb fragment cloned from C. tyrobutyricum ...... 85 3.5 Homologous recombination of pPtbEm containing an Emr-inserted ptb fragment with the chromosome of wild type resulting in the disruption of the ptb gene ...... 86 3.6 PCR characterization of the ptb mutant. A) using ptb primers: B) using Emr primers ...... 87 3.7 Comparison of activities of PTA, PTB, AK, and BK in wild type(wt) and ptb mutant ...... 88 3.8 Effect of ptb disruption on B/A ratio in batch fermentation of glucose by C. tyrobutyricum wild type and ptb mutants ...... 89 3.9 Noncompetitive inhibition of butyric acid on of C. tyrobutyricum wild type(wt) and ptb mutant (ptb) ...... 90 4.1 Construction of fibrous-bed bioreactor ...... 110

xv 4.2 Results of factor screening by DOE ...... 118 4.3 Comparison of the specific growth rates, butyric acid yields, and hydrogen yields from the simplified medium and the original CGM medium ...... 119 4.4 Cell growth and organic acid production from different substrates ...... 120 4.5 Effect of initial glucose concentration on cell growth ...... 121 4.6 Effect of pH on cell growth ...... 122 4.7 Fermentation kinetics of glucose by wild type and ptb mutant ...... 123 4.8 Fermentation kinetics of xylose by wild type and ptb mutant ...... 124 4.9 Fermentation kinetics of fed-batch free-cell fermentation of glucose by the ptb mutant at 37°C, pH 6.0 ...... 125 4.10 Fermentation kinetics of fed-batch fermentation by ptb mutant from glucose in the FBB at 37°C, pH 6.0 ...... 126 5.1 The enzymes involved in the pathway from Pyruvate to the end products ...... 152 5.2 The genomic DNA quality check for 454 sequencing ...... 153 5.3 Structure of plasmid pCAAD ...... 154 5.4 Construction of plasmid pSOS-adhE2 ...... 155 5.5 The restriction endonuclease system check with protoplast and cell crude extract ...... 156 5.6 Growth kinetics of Clostridium tyrobutyricum wild type and mutants ACK, PTA, and HydEm under various amounts of butyrate at pH 6 ...... 157 5.7 Growth kinetics of Clostridium tyrobutyricum wild type and mutants PTA, PAK, and HydEm under various amounts of butanol at pH 6 ...... 158 5.8 Comparison of specific growth rate under various amount of butanol between WT and mutant of C. tyrobutyricum ...... 159 5.9 Butanol production of C. tyrobutyricum (pCAAD) and wild type in RCM and P2 medium in serum bottles ...... 160

xvi 5.10 Free-cell fermentation kinetics of C. tyrobutyricum (pCAAD) at 30°C, pH 4.5 ...... 161 A.1 White/blue screening for gene library construction ...... 201 A.2 Random quality check of plasmids ...... 202 A.3 Random quality check of PCR product...... 202 A.4 Random quality check of purification ...... 202 A.5 Photograph of printed microarray slide and the scanned results after hybridization with cy3-labeled oligomers ...... 203 A.6 Scheme of shotgun DNA microarray construction ...... 204 B.1 Genes annotated in the 454 sequencing database for the glycolysis pathway of C. tyrobutyricum ...... 206 B.2 Alignment of two CoA transferase in C. tyrobutyricum ...... 207 C.1 HPLC chromatogram for standard sample containing glucose, xylose, lactose, acetic acid, lactic acid and butyric acid (2g/l each) ...... 212 C.2 GC chromatogram for the standard sample containing

hydrogen (80%) and CO2 (20%) ...... 213 C.3 GC chromatogram for standard sample containing acetone, ethanol, butanol, acetic acid and butyric acid ...... 214

xvii List of Tables Table Page 3.1 Strains and plasmids used in this study...... 81 4.1 Experimental design and responses for screening medium components to simplify the medium composition ...... 114 4.2 Comparison of product yields, productivities and specific growth rates of wildtype and ptb mutants in batch fermentation with glucose and xylose ...... 115 4.3 Comparison of product yields, final product concentration, B/A ratio from free-cell and FBB fermentation by ptb mutant in ...... 116 5.1 Bacterial strains and plasmids used in this study ...... 149 5.2 Location of genes and consensus % against ATCC824 ...... 150 5.3 Butanol and ethanol production by C. tyrobutyricum wild type and mutant pCAAD and pSOS-adhe2 grown in RCM and P2 medium in serum bottles at 37°C, initial pH 6 ...... 151 B.1 Location of genes involved in glycolysis in the 454 sequencing database of C. tyrobutyricum ...... 208 B.2 Comparison of amino acids sequences of two CoA from C. tyrobutyricum against other bacteria ...... 209

xviii

CHAPTER 1

INTRODUCTION

Clostridium tyrobutyricum is a member of clostridia family and was discovered

as cause for the late-blowing defect in cheese. It produces butyric acid, acetic acid, and

hydrogen from diverse carbon sources including glucose and xylose (Klijn et al. 1995).

Butyric acid is a 4-carbon and has many applications in food, perfume

and pharmaceutical industries (Dziedzak 1986; Pouillart 1998; Williams et al. 2003). The

production of butyric acid from sustainable biomass has become a promising alternative

to the current petroleum chemical route due to public concerns about environmental

pollution and the depletion of petroleum resources. Compared with other bacteria in

butyric acid fermentation, C. tyrobutyricum has many advantages, such as high product

purity and yield (Wu and Yang 2003). Furthermore, hydrogen as the gas by-product is a

clean, efficient and sustainable energy source with the highest energy content per unit

weight (143 GJ/ton) among all known fuels (Boyles 1984). In addition, hydrogen does

not contribute to any environmental problems since its oxidation product is H2O only.

The suggested metabolic pathway in C. tyrobutyricum is shown in Figure 1.1. In this pathway, the substrate, hexose or pentose, is oxidized to acetyl-CoA, with the releasing of CO2 and hydrogen. Then, acetyl-CoA can be either converted to acetyl

1

phosphate resulting in the excretion of acetate, or condensed and reduced to form butyryl-

CoA resulting in the excretion of butyrate. Phosphotransacetylase (PTA) and acetate kinase (AK) catalyze the formation of acetate and generate ATP for cell growth (Green et al. 1996). Mutants with disrupted pta and ack have been developed, and both of them showed increased butyrate and decreased acetate production (Zhu et al. 2004; Liu et al.

2006; Liu and Yang 2006).

The butyrate formation pathway comprising of two consecutive enzymatic reactions catalyzed by phosphotransbutyrylase (PTB) and butyrate kinase (BK), encoded by ptb and buk, respectively (Walter et al. 1993), plays a critical role in butyric acid

biosynthesis in C. acetobutylicum (Green et al. 1996; Green and Bennet 1997).

(1) butyryl-CoA + Pi → Butyryl-P + CoA

(2) butyryl-P + ADP → Butyrate + ATP

This pathway is important in the energy metabolism of the organism, as ATP is

produced during the conversion of butyryl-CoA to butyrate (Valentine and Wolfe 1960).

Both enzymes have been purified and characterized from C. acetobutylicum strain ATCC

824. PTB is an octamer of identical 31-kDa subunits consisting of 302 residues and BK

was reported to be a dimer of identical 30-kDa subunits consisting of 356 residues. ptb

and buk are located in the same operon with ptb proceeding buk 44 bp in C.

acetobutylicum (Hartimanis and Gatenbeck 1984; Wisenborn et al. 1989). buk has been

interrupted in C. acetobutylicum and the mutant showed a decreased butyrate production

and increased acetate production (Desai et al. 1999). The butyrate formation consumes

2

NADH, in which the reduction equivalence could be transferred to ferredoxin by NADH-

ferredoxin , then the reduced ferredoxin could be oxidized and the hydrogen is released under the of hydrogenase.

The main problem associated with conventional carboxylic acid fermentation is the final product inhibition, which causes low product concentration, yield and productivity, resulting in a high cost in product recovery (Michel-Savin et al. 1990a;

Michel-Savin et al. 1990b; Michel-Savin et al. 1990c). This problem can be partially addressed by using the fibrous-bed bioreactor in which cells are immobilized in a fibrous matrix to achieve a high cell density (Yang 1996). The fibrous-bed bioreactor has been widely applied in the production of organic acids, including propionic acid, lactic acid and acetic acid, to increase cell density, reactor productivity, titer and yield (Yang et al.

1994; Silva and Yang 1995; Huang et al. 1998). The fibrous-bed bioreactor has also been used to adapt C. tyrobutyricum to increase its butyrate tolerance and to enhance the final product concentration in the fermenter (Zhu et al. 2004; Liu et al. 2006).

Genome sequencing discovered that C. tyrobutyricum has butanol dehydrogenase and it could produce butanol if the butyryl aldehyde dehydrogenase is functional in the cell. Moreover, pta and ack mutant have shown higher butanol tolerance (16 g/l) than wild type strain (12 g/l), which is similar to C. acetobutylicum. These two facts make it rational to construct C. tyrobutyricum mutants as novel butanol producing bacteria with higher butanol tolerance.

1.1 Objectives

3

The overall goal of this research was to explore the potential of Clostridium

tyrobutyricum for production of biofuels and bio-based chemicals. First of all, a ptb-

disrupted mutant was constructed and characterized in order to evaluate the effect of ptb

knock-out on the fermentation. Secondly, the fermentation kinetics was compared

between the wild type and ptb mutant. In addition, several fermentation modes were

investigated. The fibrous bed bioreactor was also used to increase the butyric acid

tolerance. The third objective of this study was to establish butanol production in C.

tyrobutyricum by heterologous gene expression. The 454 genome sequencing study

suggested that C. tyrobutyricum can be a good butanol producing bacterium if butyryl

CoA can be reduced to butyrylaldehyde in the cell. Therefore, efforts were put on the expression of aldehyde dehydrogenase in C. tyrobutyricum. Figure 1.2 provides an

overview of the research objectives, approaches, and the scope of this study, which is

briefly described below.

Task 1: Construction and characterization of ptb-disrupted mutant by homologous

recombination

Gene disruption has been widely used in the study of protein function and gene

expression regulation. An antibiotic resistant gene (Emr) was inserted into the target gene

fragment (ptb), resulting in the target gene interruption. However, the genome of C. tyrobutyricum has not been completely sequenced and its ptb gene sequence was not available. In this case, the amino acid sequence in the highly conserved region of PTB

4

and the Clostridium preferred codon were used to design degenerate primers to achieve

the aim. After the mutant was proved by PCR identification, the related acid forming

enzyme activities were evaluated. The butyrate tolerance and the fermentation profile

were also compared between the wild type and the ptb-disrupted mutant. The results are

reported in Chapter 3.

Task 2: Fermentation kinetics and process development for butyric acid and

hydrogen production in ptb mutant.

The effects of ptb-disrupted on cell growth as well as butyric acid, acetic acid, and hydrogen production from glucose and xylose in free-cell batch fermentations were studied and the results are reported in chapter 4. In addition, medium composition was simplified and different low cost substrates were evaluated. The fermentation conditions were optimized for the ptb mutant fermentation. The fibrous-bed bioreactor was used to adapt cells for increased tolerance to butyric acid and to attain high final butyric acid concentration in the fermenter. The potential of using ptb-disrupted mutant for butyrate

and hydrogen production are also discussed.

Task 3: Establishment of butanol production in Clostridium tyrobutyricum.

The genome of C. tyrobutyricum has been sequenced by using 454 technology.

The genes encoding the key enzymes in the suggested metabolic pathway were located in

5 the database and the homology was compared with other Clostridium species.

Expression plasmid pCAAD and pSOS-adhE2 containing aad and adhE2, respectively, were constructed and transformed into C. tyrobutyricum. Fermentations were carried out in serum bottles to verify butanol production in the mutants. Free cell fermentation in a stirred-tank fermenter was also performed to evaluate the kinetics of the mutant, and the results are reported in Chapter 5.

6

1.2 References

Boyles, D. (1984). Bioenergy technology thermodynamics and costs. New York, Wiley.

Desai, R. P.;L. M. Harris;N. E. WelkerE. T. Papoutsakis (1999). "Metabolic flux analysis elucidates the importance of the acid-formation pathways in regulating solvent production by Clostridium acetobutylicum." Metabo. Eng. 1: 206-213.

Dziedzak, J. D. (1986). "Biotechnology and flavor development: An industrial research perspective " Food Technology 40: 8-20.

Green, E. M.G. N. Bennet (1997). "Genetic manipulation of acid and solvent formation in Clostridium acetobutylicum ATCC824." Biotech. Bioeng. 58: 215-221.

Green, E. M.;Z. L. Boynton;L. M. Harris;F. B. Rudolph;E. T. PapoutsakisG. N. Bennett (1996). "Genetic manipulation of acid formation pathways by gene inactivation in Clostridium acetobutylicum ATCC824." Microbiol 142: 2079-2086.

Hartimanis, M. G.S. Gatenbeck (1984). "Intermediary metabolism in clostridium acetobutylicum: levels of enzymes envolved in the fermentation of acetate and butyrate." Appl. Environ 47: 1277-1283.

Huang, Y. L.;K. Mann;J. M. NovakS. T. Yang (1998). "Acetic acid production from fructose by Clostridium formicoaceticum immobilized in a fibrous-bed bioreactor." Biotech. Prog. 14: 800-806.

Klijn, N.;F. Nieuwenhof;J. Hoolwerf;V. d. W. C.A. Weerkamp (1995). "Identification of Clostridium tyrobutyricum as the causative agent of late blowing in cheese by species- specific PCR amplification." Appl. and Enviro. Microbiol. 61(8): 2919-2924.

Liu, X.;Y. ZhuS. T. Yang (2006). "Construction and characterization of ack deleted mutant of Clostridium tyrobutyricum for enhanced butyric acid and hydrogen production." Biotech. Prog. 22: 1265-1275.

Liu, X. G.S. T. Yang (2006). "Kinetics of butyric acid fermentation of glucose and xylose by Clostridium tyrobutyricum wild type and mutant." Proc. Biochem. 41(4): 801-808.

Michel-Savin, D.;R. MarchalJ. P. Vandecasteele (1990a). "Butyrate production in continuous culture of Clostridium tyrobutyricum: effect of end-product inhibition." Appl. Microbio. Biotech. 33(2): 127-131.

7

Michel-Savin, D.;R. MarchalJ. P. Vandecasteele (1990b). "Control of the selectivity of butyric acid production and improvement of fermentation performance with Clostridium tyrobutyricum." Appl. Microbio. Biotech. 32(4): 387-392.

Michel-Savin, D.;R. MarchalP. Vandecasteele (1990c). "Butyric fermentation; Metabolic behavior and production performance of Clostridium tyrobutyricum in a continuous culture with cell recycle." Appl. Microbiol. Biotechnol. 34(172-177).

Pouillart, P. R. (1998). "Role of butyric acid and its derivatives in the treatment of and homoglobinopathies." Life Sci. 63(20): 1739-1760.

Silva, E. M.S. T. Yang (1995). "Kinetics and stability of a fibrous-bed bioreactor for continuous production of lactic acid from unsupplemented acid whey." J. Biotechnol. 41: 59-70.

Valentine, R. C.R. S. Wolfe (1960). "Purification and role of phosphotransbutyrylase." J. Bio. Chem. 235: 1948-1952.

Walter, K.;R. Nair;J. CaryG. Bennett (1993). "Sequence and arrangement of two genes of the butyrate-synthesis pathway of Clostridium acetobutylicum ATCC824." Gene 134: 107-111.

Williams, E. A.;J. M. CoxheadJ. C. Mathers (2003). "Anti-cancer effects of butyrate: use of microarray technology to investigate mechanisms." Proc. Nutr. Soc. 62: 107-115.

Wisenborn, D. P.;F. B. RudolphE. T. Papoutsakis (1989). "Phosphotransbutyrylase from Clostridium acetobutylicum ATCC824 and its role in acidogenesis." Appl. and Enviro. Microbiol. 66: 317-322.

Wu, Z. T.S. T. Yang (2003). "Extractive fermentation for butyric acid production from glucose by Clostridium tyrobutyricum." Biotech. Bioeng. 82: 93-102.

Yang, S. T. (1996). Extractive fermentation using convoluted fibrous bed bioreactor. U.S. Patent NO. 5563069.

Yang, S. T.;H. Zhu;Y. LiG. Hong (1994). "Continuous propionate production from whey permeate using a novel fibrous bed bioreactor." Biotech. Bioeng. 43(11): 1124-1130.

Zhu, Y.;X. LiuS. T. Yang (2004). "Construction and characterization of pta gene-deleted mutant of Clostridium tyrobutyricum for enhanced butyric acid fermentation." Biotech. Bioeng. 90: 154-166.

8

Figure 1.1 The metabolic pathway for the biosynthesis of butyrate, acetate, and hydrogen in Clostridium tyrobutyricum. (pfor: pyruvate: ferredoxin oxidoreductase; nfor: NADH:ferredoxin oxidoreductase; hyd: hydrogenase; pta: phosphotransacetylase; ack: acetae kinase; hbd: 3-hydroxybutyryryl CoA dehydrogenase; ptb: phosphotransbutyrylase; buk: butyrate kinase)

9

Figure 1.2 Research objectives and scope of this study

10

CHAPTER 2

LITERATURE REVIEW

2.1 Bio-based chemicals production

2.1.1 Biomass

The depletion of petroleum resource is aggravated by the increasing demand for

fossil fuels. The gasoline price per gallon increased to $ 4.10 in 2008 from $2.56 in 2005

and $3.23 in 2006 (News). Although the gasoline price has decreased in 2009, the

fluctuating and unstable oil price brings concerns on both economies and politics.

Moreover, the burning process of fossil fuel produces pollutants, causing global warming

and acid rain (Yat et al. 2008). These concerns promote intensive efforts on the

development of economical processes for producing fuels and chemicals from renewable

resources (Huber et al. 2006). In this respect, plant biomass is the most sustainable

feedstock for production of organic chemicals and biofuels (Klass 1998; Kumar et al.

2009).

The United States consumed ~25% of global energy and was responsible for ~25%

of global CO2 emissions with only 4.5% of the world’s population (Yat et al. 2008). It has been estimated that the U.S. could produce 1.3 x 109 metric tons of dry biomass per

year (Perlack et al. 2005). This amount of biomass has the energy content of 3.8×109 boe

(barrels of oil energy equivalent) which will account for around 52% of energy demand

11

while the U.S. consumes 7.3×109 barrels of oil (Book CIA)[7]. The worldwide raw

biomass energy potential in 2050 is estimated to be between 25×109 to 76×109 boe (Haug

2004). These data indicate the great potential application of biomass in the energy

industry and the chemical industry.

Plant biomass is one of the most abundant renewable resources and constantly

being formed by the interaction of CO2 in the air, water, and sunlight by plant during the

process of photosynthesis (Kumar et al. 2009). It consists of lignin and polysaccharides,

including starch, cellulose and hemicellulose. Lignocellulosic materials including

agricultural residues (corn stover, wheat straw, cane bagasse,), forest products (hardwood

and softwood), and dedicated crops (switchgrass, salix) are renewable sources of energy.

Lignin in lignocelluloses forms a protective barrier preventing the cellulose and

hemicellulose accessible to fungi and bacteria for hydrolysis. In order to convert the

biomass to fermentable sugars, plant biomass needs to be pretreated to remove the lignin

and make the cellulose in the plant fibers exposed for hydrolysis. Pretreatment techniques

include ammonia fiber explosion, chemical treatment, biological treatment, and steam

explosion (Hsu et al. 1980; Broder et al. 1995). Then, acids or enzymes can be used to

break down the cellulose or hemicellulose into its monomer as fermentable substrate.

Some microorganisms even can utilize the hemicellulose or cellulose directly. For

example, Clostridium thermocelum has the ability to convert cellulose to ethanol without hydrolysis. Three major hydrolysis processes are typically used to produce a variety of sugars suitable for ethanol production: dilute acid, concentrated acid and enzymatic hydrolysis (Broder et al. 1995). In a word, the polysaccharide in plant biomass could be

12

hydrolyzed to glucose, xylose and other fermentable sugar, then the hydrolytes are used

as feedstock for biological fermentation to produce biofuels or bio-based chemicals

(Mosier et al. 2005).

2.1.2 Bio-based chemicals

2.1.2.1 Hydrogen

Hydrogen is an efficient and clean energy source for the future. It has the highest

energy content per unit weight as high as 143 GJ/ton among all known fuels (Boyles

1984). Not chemically bound to carbon and other elements, the only oxidation product of

hydrogen is H2O. Therefore, it doesn’t contribute to the greenhouse effect and pollution.

Moreover, hydrogen can be easily converted to electricity by fuel cell to overcome the

problems caused by inconvenient transportation and storage.

Currently, 95% of hydrogen is produced through water gas shift reaction, steam

reforming of natural gas, and electrolysis of water (Elam et al. 2003). However, all of these methods either use fossil fuel as materials, or are highly energy intensive. The alternative method to those chemical routes is biological production, which includes direct or indirect biophotolysis, fermentation, photosynthetic production and in vitro enzymatic conversion of biomass (Woodward et al. 1996). Among these biological routes, fermentation is one of the most promising technologies contributed to the following advantages: (1) it takes renewable biomass as substrate; (2) it produces valuable metabolites as byproducts; (3) the bioreactor design is simple for the dark fermentation (Nath and Das 2004). Many microorganisms, such as Clostridium,

13

Enterobacter, Rhodopseudomonas, etc., have the ability to produce hydrogen during

anaerobic fermentation process via reactions catalyzed by iron- or nickel-containing

hydrogenases (Kataoka et al. 1997; Kumar et al. 2001; Zhu et al. 2004; Liu et al. 2006;

Liu and Yang 2006). These hydrogenases catalyze the reversible reaction between

molecular hydrogen and its component two protons and two electrons.

+ -1 2H +2e → H2

Generally molecular hydrogen is formed during the transfer of reducing

equivalents produced during pyruvate oxidation using pyruvate-ferredoxin

oxidoreductase (PFOR) coupling with hydrogenase (Gray and Gest 1975; Cammack et al.

2001; Hallenbeck and Benemann 2002). The reactions are illustrated by the following

equations:

Pyruvate + CoA + ferredoxin (ox) → Acetyl-CoA + CO2+ ferredoxin (red)

Ferredoxin (red) → Ferredoxin (ox) + H2

The theoretical molar yield of hydrogen is two per mol glucose if the reducing

equivalent from pyruvate is the only reducing source of the hydrogen. However,

Clostridium also contains NADH: ferredoxin oxidoreductase (NFPR) which can produce

additional hydrogen using the reducing equivalent stored in NADH during the glycolysis

(Jungermann et al. 1971; Hallenbeck 2005).

NADH + ferredoxin (ox) → NAD++ ferredoxin (red)

Ferredoxin (red) → Ferredoxin (ox) + H2

The oxidation of NADH by ferredoxin (ox) is energetically unfavorable under

standard condition and ΔG could be negative only at low partial pressures of hydrogen

14

(Hallenbeck 2005; Kraemer and Bagley 2007). This challenge could be addressed by

coupling with the ATP hydrolysis which is highly energetically favorable.

The key issue for the fermentative hydrogen production is the low yield. The

maximum hydrogen yield from Clostridium is 4 if the acetate is the only end-product.

However, if butyrate, which consumes 2 mol of NADH during its formation, is the end- product, only 2 mol of hydrogen can be achieved (Thauer et al. 1977). It has been

estimated the overall cost of hydrogen production using fermentable biomass as

feedstock is $25/GJ based on some assumptions (Classen et al. 2000). Therefore, the

hydrogen is currently more expensive than other alternative fuels. NREL conducted an

economic analysis indicating that the a hydrogen price could hit $2.47 kg-1 which is in the

range of the hydrogen cost goal with assumptions of that the glucose price is under 0.11

kg-1 and the molar yield of hydrogen is 4 mol per mol glucose (Turner et al. 2008).

Both molecular manipulation and process development have been used to

improve hydrogen production. The anaerobic fermentation leads to production of volatile

fatty acids and . The production of alcohol and reducing acid such as lactic acid

consumes the reducing equivalent (NADH) and causes low hydrogen production (Levin

et al. 2004). In order to enhance the hydrogen yield, the carbon source needs to be

directed to the acetate pathway, instead of alcohol, lactic acid and other volatile fatty acid

pathways (Hawkes et al. 2002; Levin et al. 2004). The amount of NADH could be

increased if the pathway to alcohol or lactic acid was blocked and the hydrogen could be

formed during the oxidation of NADH (Das and Veziroglu 2001). Kumar et al. (Kumar et

al. 2001) used the proton-suicide technique with NaBr and NaBrO3 to block the pathways

15

of organic acid formation and increased hydrogen production to 3.8 mol/mol glucose. A

similar enhancement of hydrogen yield was achieved by blocking the biosynthesis of

alcoholic and acidic metabolites using both ally alcohol and the proton suicide technique

in E. aerogenes HU-101 (Mahyudin et al. 1997). Gene inactivation has also been employed to improve the hydrogen molar yield from 0.56 to 1.17 in Enterobacter aerogenes (Rachman et al. 1997). Hydrogenase is another target for hydrogen production. Some hydrogenases interact with NADH in the cytoplasmic side and with proton in the periplasmic side to form hydrogen. The reverse transcription PCR and quantitative real-time PCR methods have been developed to investigate the transcription level of hydA during hydrogen fermentation by C. butyricum (Tolvanen et al. 2008;

Wang et al. 2008). Wang et al. showed linear correlation between the specific hydrogen production rate and the expression level of hydA (hydrogenase A) (Wang et al. 2008). It is also noticed that the profiles of bacterial growth displayed a similar trend to the hydA expression level.

On the other hand, fermentation process conditions were optimized to increase the bio-hydrogen production. The effect of hydraulic retention time on the hydrogen yield was studied (Chu et al. 2009). It is concluded that at the optimum HRT range 6–8 h, the hydrogen yield was in the range of 1.4–1.5 mol H2/mol glucose. Cheng et al. (Cheng et

al. 2002a) reported that pH condition of inoculum (sewage-treated plant sludge)

enhanced hydrogen production and shortened the lag time. A dual-substrate steady-state

model has been developed and the model predicted the relation between the accumulation

of the end product and the consumption of glucose and peptone at the dilution rate 0.06-

16

0.8 per hour (Whang et al. 2006). Other factors such as pH, heat shock, the concentration

of substrate, the concentration of Cu2+, and the concentration of Fe2+ have also been

explored (Lee et al. 2001; Cheng et al. 2002b; Lee et al. 2002; Noike et al. 2002; Bai et al. 2004; Lee 2004).

2.1.2.2 Butyric acid

Butyric acid is a 4-carbon carboxylic acid and has many applications in food, perfume and pharmaceutical industries. Its acid form could enhance the butter-like flavor in food and the esters of butyric acid serve as aromatic compounds in perfume industries

(Dziedzak 1986). It is also used as a raw material for the production of biodegradable polymer β–hydroxylbutyrate. Butyrate has been used as the building blocks to synthesize the plastic material and textile fiber. Moreover, butyrate is considered as an anti- neoplastic drug because of its ability to inhibit the histone deacetylases (HDAC) and the favorable safety profile in humans (Pouillart 1998; Williams et al. 2003). A family of acyloxylalkyl butyrate prodrugs is presently in clinical development (Rephaeli et al.

2000). Butyric acid derivatives have also been developed to produce antithyroid and vasoconstrictor drugs and used in anesthetics (Playne 1985).

Currently, butyric acid is produced by oxidation of butyraldehyde obtained from oxosynthesis of propylene (Pryde 1978; Kroschwitz and Howe-Grant 1997). The production of butyric acid from fermentation of biomass has become an promising alternative to the current petroleum-chemical synthesis. Butyric acid producing bacteria, such as C. tyrobutyricum, C. butyricum, C. beijerinckii, C. populeti, C. barkeri, and C.

17

thermobutyricum produce butyric acid as their main product from various substrates and

also form acetate as byproduct (Crabbendam et al. 1985; Michel-Savin et al. 1990a;

Michel-Savin et al. 1990b; Michel-Savin et al. 1990c) (Alam et al. 1988; Patel and

Agnew 1988; Zigova and Sturdisk 2000). Compared to other bacteria in butyric acid

fermentation, C. tyrobutyricum has many advantages such as high product purity and

yield (Wu and Yang 2003). C. tyrobutyricum was discovered as notorious bacteria for its

ability to ferment lactate in the cheese to produce gas and high level of butyric acid,

which causes the blowing of the cheese. The early researches were focused on how to

identify and limit C. tyrobutyricum (Klijn et al. 1995).

In order to improve the production of butyric acid, a fibrous-bed bioreactor (FBB)

was successfully used for butyrate fermentation. The reactor productivity and final

product concentration was increased (Zhu and Yang 2003). With the immobilization of

cells in the fibrous matrix, high cell density was attained and the reactor productivity,

final product concentration, and product yields were improved. The butyrate yield

achieved 0.423 g/g or 0.9 mol/mol of glucose consumed in the fermentation (Zhu et al.

2002). Molecular methods were also employed to improve strain and direct the carbon

source to butyric acid formation. pta and ack gene encoding PTA and ACK were

disrupted by the insertion of erythromycin resistant gene (emr) and the mutant shows

higher butyric acid production than wild type (Zhu et al. 2004; Liu and Yang 2006). Fed

batch fermentation was conducted to examine the fermentation kinetics of the mutants.

The butyrate final concentration reached 32.5 g/l and 41.6 g/l, for pta mutant and ack

mutant respectively, much higher than the 20.2 g/l for wild type. The fibrous bed

18 bioreactor was also performed to adapt cells under a high butyric acid concentration to increase the final titer (Zhu et al. 2002; Liu and Yang 2006). Several bio-waste such as corn fiber hydrolysate, corn meal hydrolysate and cane molasses were also used as substrate to produce butyric acid (Huang 2002; Zhu et al. 2002; Jiang et al. 2009). The effect of fermentation pH, the substrate, and the initial concentration of substrate on the fermentation profile has been evaluated (Zhu et al. 2004; Jo et al. 2008). Similar to other fermentation process suffering from product inhibition, C. tyrobutyricum fermentation is strongly inhibited by butyric acid (Michel-Savin et al. 1990a). Researcher developed in situ separation processes to decrease the acid product concentration during the fermentation

(Keshav et al. 2009). Micro-filtration, ultra-filtration, and permeate electrodialysis have also been integrated with fermentation of lactic acid (Boyaval et al. 1987). The extractive and pertractive fermentation processes using Hostarex (20% w/w) in oleyl alcohol as the organic phase to produce butyric acid by C. butyricum have been developed. The experimental results showed that the integration of extraction and pertraction with fermentation improved the butyric acid concentration from 7.3 g/l to 10.0 g/l (extractive fermentation) and 20.0 g/l (pertractive fermentation) and that the butyric acid production yield increased from 0.24 g/g to 0.30 g/g (pertractive fermentation) (Zigova et al. 1999).

A novel extractive fermentation process by FBB was developed by Wu and Yang (Wu and Yang 2003). 10% (v/v) alamine 336 in oleyl alcohol was used as extractant in a hollow-fiber membrane extractor to selectively remove butyric acid from fermentation broth.

19

2.1.2.3 Butanol

The acetone-butanol-ethanol (ABE) fermentation was one of the oldest and

largest biological fermentation processes in the history, only second to the ethanol

fermentation. It was discovered by Louis Pasteur in 1861 (Jones and Woods 1986) and

attracted intense attention due to its ability to produce acetone which was highly

demanded as the solvent in cordite production during the WWI and the WWII (Killeffer

1927). However, the development of petroleum industry lower the production cost of

butanol chemically through oxo process(Lemke 1963) or aldol process (Maiorov et al.

1971), which caused the decline of the ABE fermentation since 1960. Recently, the rising

price and the limited resource of the petroleum trigger the renascence of the ABE

fermentation since butanol is considered as one of the most promising alternatives to

gasoline for several reasons. With longer carbohydrate chain, butanol has lower vapor

pressure (0.33 psi) (Minteer 2006) and is much safer and more environmentally friendly.

With a higher energy content than ethanol, closer octane number to gasoline and lower

affinity to water, butanol can be blended into fuels or directly replace gasoline and act as

an alternative fuel in engines (Alasfour 1997).

ABE fermentation is conducted by strictly anaerobic Clostridia. However, due to

the complicated origin of the microorganisms and the unsystematic nomenclature based on the differences in the type and the ratio of the solvent produced, the name of different clostridia species led confusion for long time (Jones and Keis 1995; Keis et al. 2001).

Keis et al. performed a system analysis based on the 16S rRNA sequence, biotyping and

DNA fingerprinting, then categorized the industrial solvent-producing clostridia into 4

20

species, which are C. acetobutylicum, C. beijerinckii, C. saccharoperbutylacetonicum sp.

nov. and C. saccharobutylicum sp. nov.. Most industrial solvent-producing clostridia

strains are classified as C. beijerinckii or C. acetobutylicum (Jones and Keis 1995; Keis et

al. 1995; Keis et al. 2001).

The ABE fermentation has two stages, acidogenesis and solventogenesis as

shown in Figure 2.1. At the exponential stage of fermentation, the Ph value decreases

while the bacteria convert saccharide to acetic acid and butyric acid. After the Ph value

lower than 4.5-5.0, the fermentation goes to stationary phase and the metabolic pathway

will be shifted to solventogenesis. The bacteria will start to produce ethanol, acetone and butanol (Jones and Woods 1986). However, in an environment containing high alcohols

like butanol, the bacterial membrane changes and the membrane becomes more fluid and

easier to be lysed (Linden and Moreira 1982). Butanol production is thus lowered since

the number of solvent producing bacteria decreases. Moreover, ABE fermentation suffers

from the low yield, low productivity, and low final concentration due to the end-product inhibition (Huang et al. 2004). Therefore, the ABE fermentation route is still not competitive with the gasoline economically. Extensive metabolic engineering studies have been done and several mutants have been constructed to increase the butanol selectivity or the butanol tolerant.

C. acetobutylicum ATCC 824, isolated from the garden soil in Connecticut in

1924 (Weyer and Rettger 1927) and closely related to the historical Weizmann strain

(Johnson and Chen 1995; Cornillot and Soucaille 1996), has been used as the genomic model to investigate the gene expression regulation network, although this strain has

21

never been used in industry. Its genome has been sequenced and annotated (Nolling et al.

2001), containing a 3.94 Mb chromosome and a 192 kb megaplasmid pSol1 where locate

genes involved in sporulation and solvent formation, such as aad, ctfA, ctfB, adc

(Cornillot et al. 1997; Tomas et al. 2003). Several genes involved in the acid production,

solvent production and sporulation have been cloned and mutants with inactivated gene

or overexpressed gene have been constructed to examine the gene’s function and to

explore gene expression regulation by microarray and/or other methods. The mutant

without the megaplasmid (M5) displays an asporogenous and cannot produce solvent.

This confirmed that the megaplasmid plays an important role in solvent production

(Tomas et al. 2003). In the butyrate kinase (buk) inactivation mutant, higher level of butyryl phosphate (BuP) is detected, and consistently, solvent is formatted earlier and the concentration is higher than wild type (Desai and Papoutsakis 1999). The results suggested that the intracellular BuP is a putative regulator of solventogenesis in C. acetobutylicum, which probably acts as a phosphodonor of transcriptional factor.

Transcriptional analysis confirmed that higher BuP level in the buk mutant related to the up-regulation of solvent formation genes (Zhao et al. 2005). In alcohol/aldehyde dehydrogenase (aad) overexpression combining antisense RNA downregulation of CoA transferase B (ctfB) mutant, aad was overexpressed and resulted in the highest reported ethanol production. On the other hand, the butyrate depletion in the fermentation suggested that butyryl CoA maybe the limiting factor on butanol production. The addition of butyric acid increasing the butanol/ethanol ratio confirmed this hypothesis

22

(Harris et al. 2000; Tummala et al. 2003). Moreover, the enzyme PTB and BK were found able to uptake butyrate without acetone production (Desai et al. 1999).

Besides the gene manipulation, fermentation condition has also been optimized

and strains have also been adapted by different methods to increase the solvent production. Cell immobilization to increase butanol production and prolong the fermentation lifetime is also widely studied(Larger et al. 1985; Lienhardt et al. 2002).

Considering the important roles of butyrate, BuP and Butyryl CoA in the solvent

production, the butyrate co-feeding with glucose strategy could probably increase butanol

yield since both BuP and ButyrylCoA concentrations would be higher during the

butyrate/glucose co-feeding. Experiments and stoichiometric modeling indicate that

butanol production can be increased by co-feeding sugar with butyric acid (Shinto et al.

2007). A two-step fermentation process has been developed (Ramey 1998) which can

increase the butanol production by feeding the C. acetobutylicum with sugar and butyric acid produced from saccharide by C. tyrobutyricum.

The effects of glucose, butyrate, pH and dilution rate on the butanol production in continuous fermentation were also investigated. A high butanol yield of 0.42 g/g glucose was obtained by continuous production in a fibrous-bed bioreactor using butyrate and glucose as cosubstrates (Huang et al. 2004).

2.2 Metabolic engineering and strain development

2.2.1 Clostridium

23

The genus Clostridium is a heterogeneous assemblage of gram-positive,

obligatory anaerobic, rod-shaped bacteria forming endospores under extreme conditions

(Young and Cole 1993). The heterogeneity of the group is indicated by the remarkable

range of d(G+C) contents of their DNAs from 24% to 55% (Johnson and Francis 1975;

Cato et al. 1986). Clostridium attracts intense interest due to its wide application in

medical research and fermentation industry. C. perferingens is the pathogenic strain

causing fatal Necrotic enteritis (pig-bel). C. tetani produces tetanospasmin, a biological

toxin, and is the causative agent of tetanus. C. thermocelum, C. stercocellulolyticum and

C. cellulovorans have the ability to degrade polysaccharide including cellulose, hemicellulose and starch in biomass. Some Clostridium strains produce organic solvents and acids through anaerobic fermentation from sugar, such as extensively investigated C. acetobutylicum, which has been well known for the Acetone-Butanol-Ethanol (ABE) fermentation. Until now, the whole genomes have been sequenced for 5 Clostridium species, and some other species sequencing projects are undergoing.

C. tyrobutyricum is a member of Clostridia family and it was discovered as the causative agent for the late-blowing defect in cheese (Klijn et al. 1995). Under anaerobic fermentation condition, it produces butyric acid, acetic acid, and hydrogen. The suggested metabolic pathway was illustrated in Figure 2.2. Compared to other Clostridia bacteria, C. tyrobutyricum does not produce any alcohol, simplifying the final product composition. It also has many advantages including simple medium composition, and relatively high product purity and yield (Wu and Yang 2003). Gene inactivation through

24 homologous recombination has been successfully applied to block the acetate pathway to increase butyrate production (Zhu et al. 2004; Liu and Yang 2006).

2.2.2 Metabolic engineering

Metabolic engineering is defined as the method of optimizing the genetic and regulatory process by recombinant DNA techniques including gene disruption, gene insertion (heterologous expression) and gene overexpression technology. The purpose of the metabolic engineering is to understand the metabolic network and to direct the flux to the desired products. It provided many useful biosynthesis tools for biotechnology applications (Michalodimitrakis and Isalan 2009). Metabolic flux analysis (MFA) has been used to analyze and predict the flux distribution in the metabolic pathway (Varma and Palsson 1994; Stephanopoulos et al. 1998). The stoichiometric models are established by a set of equations using metabolic flux analysis (MFA), where the fluxes are measured by experiment. Linear programming was used as a major tool to develop the MFA. Genome scale models of microbial cells have been developed to predict the impact of a metabolic pathway on cell growth and product biosynthesis

(Price et al. 2004). Kayser et al. (Kayser et al. 2005) studied the native metabolism of E. coli under different growth conditions while Ozken (Ozkan et al. 2005) did a similar research for recombinant protein production.

Metabolic engineering has been widely applied to the production of bio-based chemicals or biofuels from renewable biomass. For example, the genes encoding enzymes responsible for biosynthesis of isopropanol (Chen and Hiu 1986; Hanai et al.

25

2007) and butanol (Atsumi et al. 2008) in Clostridium were recently expressed and the

production of isopropanol or butanol was established in E. coli. The pathway for higher

alcohol production from amino acid by S. cereviciae (Sentheshanmuganathan 1960;

Yoshimoto et al. 2002; Schoondermark-Stolk et al. 2006) was expressed in E. coli. Six

straight and branched-chain alcohols were produced and 1.28 g/l of isopentanol have

been produced by directing the flux to the desired pathway (Kalscheuer et al. 2006;

Ladygina et al. 2006; Connor and Liao 2008).

Other method for metabolic engineering is to adjust the redox balance. The

availability of NADH or NADPH could be increased or NADH and NADPH could be

switched by overexpression of corresponding enzymes (Dos Santos et al. 2004; Sanchez

et al. 2006). The expression of heterologous gene gyceraldehydes-3-phosphate

dehydrogenase results in a 40% reduction in glycerol production and a 3% increase in

ethanol production when the gene was introduced into S. cerevisiae (Bro et al. 2004; Bro

et al. 2006). The deletion of NADPH dependent glutamate dehydrogenase led an altered

redox metabolism and hence changed expression of other genes encoding NADPH-

dependent enzymes. Hydrogen production from glucose was increased by blocking

alcohol and some organic acid formation pathways in E. cloacae IIT-BT-08 and

hydrogen yield was enhanced by 62% in E. cloacae double mutant than the wild type

strain (Kumar et al. 2001).

Instead of one-gene-a-time, combinatorial engineering was developed to express multiple gene (Pfleger et al. 2006) in one mutant. Some other techniques such as ligation-

26

free assembly (Li and Elledge 2007) and BioBricks (Shetty et al. 2008) could construct

pathways from existing DNA fragments or genes.

Introducing genes for uptaking different substrates into cells has also been studied.

pUR400 plasmid carrying invertase gene, responsible for breaking down sucrose to

glucose and fructose, is transformed in to E.coli strain HD701 and successfully utilize sucrose to produce hydrogen (Penfold et al. 2003).

Hydrogenase is another target for hydrogen production. Karube (Karube et al.

1983) cloned and expressed the hydrogenase gene from C. butyricum in E. coli strain

HK16 (hyd-) in CB medium resulting that the hydrogenase activity increased by 3 folds

compared to wild type (Karube et al. 1983). hydA, a [Fe]-hydrogenase gene from E.

cloacae IIT-BT-08 was overexpressed E. coli BL-21 which does not produce hydrogen,

the hydrogen yield was enhanced (Mishra et al. 2004; Chittibabu et al. 2006) to 3.12,

higher than in the wild type IIT-BT-08. Yoshida and colleagues blocked the lactate (ldhA

knock-out) and succinate (frdBC knock-out) production pathways, the hydrogen yield

was increased by 90% (Yoshida et al. 2006). Multiple knock-out/overexpression E. coli

strain has been constructed by the combination of fhlA overexpressed, hycA inactivated,

ldhA and frdBC deleted , the effect of those mutation were evaluated step by step

(Yoshida et al. 2006; Yoshida et al. 2007).

2.2.3 Key enzymes related to the study

The anaerobic fermentation pathway in C. tyrobutyricum is shown in Figure 2.2.

The substrate, hexose or pentose, is firstly catabolized to pyruvate via EMP pathway or

27

HMP pathway, and then oxidized to acetyl-CoA. The second oxidation process is

accompanied by ferredoxin reduction. The resulting FdH2 is then oxidized and the

electrons are transferred to proton to produce hydrogen catalyzed by hydrogenase

(Hallenbeck and Benemann 2002).

The acetyl-CoA is the branch point intermediate in the metabolic pathway, which can be converted to either acetyl phosphate resulting in the excretion of acetate, or

butyryl-CoA, which could be converted to butyryl phosphate resulting in the excretion of

butyrate. The following enzymes play critical roles in this anaerobic pathway.

2.2.3.1 PTB and BK

The butyrate formation is comprised of two consecutive enzymatic steps:

(1) butyryl-CoA + Pi → Butyryl-P + CoA

(2) butyryl-P + ADP → Butyrate + ATP

These two steps are catalyzed by phosphotransbutyrylase (PTB) and butyrate

kinase (BK), encoded by ptb and buk, respectively (Walter et al. 1993). This pathway is

important in the energy metabolism of the organism, as ATP is produced during the

conversion of butyryl-CoA to butyrate (Valentine and Wolfe 1960). Both enzymes have

been purified and characterized from C. acetobutylicum strain ATCC 824. PTB is an

octomer of identical 31-kDa subunits consisting of 302 residues and BK was reported to

be a dimer of identical 30-kDa subunits consisting of 356 residues. ptb and buk are

located in the same operon with ptb proceeding buk 44 bp in C. acetobutylicum

(Hartimanis and Gatenbeck 1984; Wisenborn et al. 1989). buk has been interrupted in C.

28

acetobutylicum and mutant shows a decreased butyrate production and increased acetate production.

2.2.3.2 PTA and AK

Similar to the PTB and BK, phosphotransacetylase (PTA) and acetate kinase

(AK) catalyze the formation of acetate and generate ATP for cell growth (Boynton et al.

1996). The pta gene has been cloned and characterized in different microorganisms,

including C. acetobutylicum (Boynton et al. 1996), E. coli (Matsuyama et al. 1989;

Kakuda et al. 1994), M. thermophila (Latimer and Fery 1993), and so on. The fragments

of these two genes in C. tyrobutyricum have also been cloned and sequenced (Zhu et al.

2004; Liu et al. 2006). The cloned pta fragment corresponding 244 amino acids and the

amino acid sequence is 70% of similarity to the PTA of C. acetobutylicum. The result

shows the high degree of similarity in PTA between C. tyrobutyricum and C.

acetobutylicum. This fragment was used to construct a pta-deleted mutant by homologous

recombination. The fermentation kinetic suggested that the pta inactivation could

enhance the butyrate productivity by 1.9-folder (Zhu et al. 2004). The ack deletion also

shows similar result.

2.2.3.3 Hydrogenase

Hydrogenases catalyze the conversion between molecular hydrogen and two

+ - protons and two electrons (H2 ↔2H + 2e ). The H2 metabolism has been studied

extensively in over sixty species. As at least 13 families of hydrogenase have been

29

identified, hydrogenases are a heterogeneous group of enzymes. Eleven out of the

thirteen families of hydrogenase are metalloenzymes containing metal centers such as

heme groups, [Fe-S] clusters, or Ni-Fe centers. The other two families contain non-metal

redox groups such as FAD and FMN. The common Ni-Fe hydrogenases consist of a

small subunit of 30 kDa and a large subunit of 60 kDa (Vignais et al. 2001; Frey 2002;

Vignais and Colbeau 2004). The most intensively studied hydrogenase is the Fe-only

hydrogenase called hydrogenase I encoded by hydA, which has been found in C.

acetobutylicum (Gorwa et al. 1996) and C. pasteurinum (Meyer and Gagnod 1991). Most of hydrogenases are oxygen-sensitive, however, Venter team discovered a novel oxygen- insensitive hydrogenase (Lee et al. 2008).

The molecular biology tools can be used to direct the carbon source and shift the

NADH to ferredoxin to increase the hydrogen production. Gene inactivation has been successful employed to improve the hydrogen molar yield from 0.56 to 1.17 in

Enterobacter aerogenes (Rachman et al. 1997). Chemical inhibitor has also been used to block the ethanol production to increase the hydrogen production in Clostridium.

2.2.3.4 Aldehyde/alcohol dehydrogenase

The solvent-producing genes in C. acetobutylicum have been cloned, sequenced and characterized in C. acetobutylicum, including aad, ctfA/B, adc encoding aldehyde/alcohol dehydrogenase, CoA transferase A/B, acetoacetate decarboxylase respectively. These genes are located in the megaplasmid pSOL1 in C. acetobutylicum.

30

However, C. tyrobutyricum does not have the megaplasmid and the ability to produce solvent. Among these genes, aad is the key gene responsible for the alcohol production.

The aad gene has 2619 base-pair and codes for a 96,517 Dalton protein. The N- terminal section of aad shows homology to aldehyde dehydrogenase of bacteria while the

C-terminal shows homology to alcohol dehydrogenase of bacteria. The entire amino acid sequence of aad exhibits 56% identity to the tri-functional protein adhE from E. coli

(Nair and Papoutsakis 1994). The aad-expressing plasmid has been introduced into C. acetobutylicum and the butanol dehydrogenase, acetaldehyde dehydrogenase and butyraldehyde dehydrogenase activity show obvious increase and ethanol dehydrogenase shows a small increase in the mutant. The same plasmid has been also transformed into the M5 mutant which does not produce solvent and the transformation restored the butanol production without any acetone and ethanol detected (Nair and Papoutsakis

1994).

Fontaine et al. (Fontaine et al. 2002) characterized the second enzyme ADHE2 which also can catalyze the reduction of aldehyde and alcohol. The enzyme was coded by adhE2 gene which has been cloned and over-expressed in the mutant DG1which is not able to produce solvent. The transformed mutant resumed butanol production and shows higher selectivity on butanol than ethanol.

2.3 Process development

2.3.1 Medium composition

2.3.1.1 Substrate selection and initial concentration

31

Unlike recombinant protein fermentation, biofuels and bio-based chemicals

fermentation values the substrate cost. Current commercial fermentation processes focus on starch based technologies. There is different opinion from both the scientific community and public arguing that the usage of food grains as feedstock does not possess environmental and economic benefits (Fargione et al. 2008). Clostridium has a broad

substrate range including glucose, xylose, lactose, starch; some species such as C.

thermocelum even can uptake cellulose directly. On the other hand, lignincellulose

consisting of lignin, cellulose and hemicellulose attracts many attentions as its low price

and sustainability. After pretreatment and hydrolysis, glucose and xylose are released and

used as fermentation substrate. Corn fiber hydrolysate has been used in butyric acid

production by C. tyrobutyricum fermentation. The medium was supplemented with corn

steep liquor as source and 0.47g/g butyric acid yield and 2.91g/l/h productivity

was attained (Zhu et al. 2002).

Meanwhile, some industrial byproducts could be used as potential substrate in the

bioprocess. Cheese whey is the byproduct from cheese production, containing whey

protein and whey lactose. Whey protein concentrations (WPCs) is extracted from the

cheese whey and the whey permeate containing lactose is left behind as byproduct, which

can be obtained at low price. Whey permeate has been used for production of bio-based

chemicals or biofuels through fermentation process.

Glycerol is another attractive substrate for fermentation process because its

availability and low price as the byproduct from biodiesel production. In addition, it is

more reduced compared to glucose or xylose and could benefit the production of

32 reductive chemicals such as hydrogen, ethanol, succinate, butanol, 1,3-propanediol (Biebl

2001; Ito et al. 2005; Dharmadi et al. 2006; Wang et al. 2007; Buelter et al. 2008).

The initial substrate concentration is another important factor in fermentation as it impacts microbial growth, yield and productivity of end products (Fabiano 2002..).

Saccharide molecules cannot enter the cell membrane freely due to its high polarity. A complicated transportation mechanism exists to facilitate the transportation of glucose and other monosaccharide (Bisson 1988; Bisson et al. 1993). S. cerevisiae has 20 genes encoding a high efficient hexose transporter for hexose (Bisson et al. 1993; Lagunas 1993;

Perez et al. 2005). Meanwhile, glucose displays different uptake mechanism depending on the glucose concentration in the environment (Fuhrmann and Volker 1992; Walsh et al. 1994). Consequently the fermentation will display different kinetics under different glucose concentration. In addition, high glucose concentration may cause the substrate inhibition. Anaerobic bacterium A. succinogenes shows tolerance to 165g/l glucose.

However, the growth rate, succinate production was significantly reduced and long lag phase was observed when the initial glucose concentration exceeds 65g/l (Guettler et al.

M. B. Institute 1996; Urbance et al. 2004). The similar inhibition pattern was observed in cane molasses fermentation (Liu et al. 2008). Lin et al. (Lin et al. 2008) developed a model to describe the glucose inhibition and confirmed the prediction using experimental data. Both model and experiment results showed a prolonged lag phase and reduction of specific growth rate with the higher initial glucose concentration (Lin et al. 2008). The main transport mechanism for carbohydrates in Clostridium involves the phosphoenolpyruvate (PEP)-dependent system (PTS) (Mitchell 1998),

33

which also takes part in gene expression and regulation (Tangney et al. 2003).It is also

proposed that the uptake of sugars appears to be driven by ion gradients (H+ or Na+ symporters) or ATP hydrolysis (ABC transporters) (Mitchell 1998).

2.3.1.2 pH effect

The pH value is another important factor for fermentation by affecting the enzyme activity via changing the ionization state of enzyme’s components (Fabiano 2002).

Consequently, pH value impacts the carbon source distribution (Zhu et al. 2004) and the product inhibition is also affected by the pH. The mechanism of short-chain organic acid inhibition was proposed as the acid form of the organic acid under the low pH in the extracellular space passes the cell membrane due to its non-polar property, then dissociated in the neutral environment in the intracellular space, the extra proton is released and causes pH drop in the intracellular space. In order to control the environment, the extra protons have to be pumped out of the intracellular space by H+-

ATPase at the price of ATP hydrolysis. In other words, under the same amount of acid radical existence, lower pH will result in more acid form and consumes more ATP to maintain the environment in the cells.

2.3.2 Fermentation mode

2.3.2.1 Immobilized cell fermentation

The main problem associated with conventional carboxylic acid fermentation is the final product inhibition, which causes low product concentration, yield and

34

productivity, resulting in high cost in product recovery and decreased cost-effectiveness

(Michel-Savin et al. 1990c). This problem can be partially addressed by the fibrous-bed

bioreactor where cells are immobilized in fibrous matrix and high cell density could be

achieved (Yang 1996). The productivity, final product concentration and production yield

are also enhanced by FBB compared to free-cell fermentation (Yang et al. 1994). FBB

fermentation has been successfully employed in the production of organic acids,

including lactic acid, propionic acid, acetic acid (Silva and Yang 1995; Huang et al.

1998; Huang and Yang 1998; Zhu and Yang 2003). It was also carried out to adapt the C.

tyrobutyricum wild type and pta/ack mutants. Both wild type and mutants were adapted

in the FBB and showed a higher tolerance to butyric acid (Zhu and Yang 2003; Zhu et al.

2004; Liu et al. 2006; Liu and Yang 2006).

Zhu and Yang discovered that the adaptation mutant by FBB showed a higher

tolerance to butyric acid and a higher H+-ATPase activity (Zhu and Yang 2003). It could

be explained by the H+-ATPase functions as a proton transporter (Riebeling and

Jungermann 1976). H+-ATPase is a membrane-bound enzyme essential for maintaining the transmembrane proton motive force required for ionic regulation and the active transport of nutrients (Riebeling and Jungermann 1976; Ivey and Ljungdahl 1986). The higher activity of H+-ATPase can assure the extra protons in the cytoplasm could be

transported to the outside of cells, maintaining proper pH in cytoplasm. Additionally, the

fatty acid composition ratio of cell membrane was changed in the adapted cells. More

saturated and long-chain fatty acids were found, resulting in less membrane fluidity (Zhu

and Yang 2003). It is consistent with the result obtained from S. cerevisiae (Casey and

35

Ingledew 1986) and E. coli (Ingram 1976). Both of them have the ability to regulate

membrane lipid composition to increase their tolerance to organic solvents.

2.3.2.2 Continuous fermentation

Generally, Continuous fermentation has several advantages compared to batch

fermentation, such as high productivity due to the continuous operation without lag phase,

resistant to the foul and contamination (Van Groenestijn et al. 2002). It can significantly increase hydrogen production (Brosseau and Zajic 1982) and butanol production (Lee et al. 2008). Magnusson demonstrated the continuous hydrogen production using cellulose as substrate by C. thermocelum (Magnusson et al. 2008). The hydrogen production rate reached 24.8 ml/l·h while Collet et al. achieved 73 ml/l·h using soluble lactose as substrate (Collet et al. 2004). Both studies showed robust of the system and easy recovery from the pH failure or other disruption. Spores may form during the failure and the fermentation experienced some recovery time (Magnusson et al. 2008). Huang et al. used butyrate co-feeding strategy by continuous fermentation and investigated the relation between the dilution rate/pH and the butanol production (Huang et al. 2004). 4.6 g/l·h productivity and 0.42 g/g yield has been achieved under the optimized conditions by C. acetobutylicum. Lee et al. (Lee et al. 2008) studied the kinetics of butanol fermentation by C. beijerinckii using batch and continuous cultures containing suspension or immobilized cells. The highest butanol productivity and yields was obtained by continuous fermentation of immobilized cells with addition of butyrate or acetate in the medium.

36

2.2.2.3 Sparging and operational controls

Maeda (Maeda et al. 2007) found that a low partial pressure system increased

hydrogen productivity in E. coli . This is probably due to the low partial pressure of hydrogen lower the Gibb’s free energy. Gas sparging usually increases hydrogen

production by decreasing the hydrogen solubility (Nath and Das 2004; Kraemer and

Bagley 2007). In addition, sufficient mixing facilitates the release of hydrogen from the

fermentation broth (Kraemer and Bagley 2007). Ceramic fittings has been used in

continuous fermentation for hydrogen production from sucrose by C. butyricum and the

hydrogen production rate achieved 307 ml/l·h without any additional energy input

(Fritsch et al. 2008). The high surface area to volume ratio can help bubbles form to

release the supersaturated hydrogen in the medium, which may induce inhibition on the

cell (Jones et al. 1999).

2.4 Genome sequence and microarray

2.4.1 Genome sequence

Whole genome sequence is critical for metabolic engineering because it can

provide the deep information of the gene and genome structure. The first whole genome

sequenced is virus φX174 in 1977 (Sanger et al. 1977). Since then, the development of

automation and the invention of PCR (Mullis et al. 1986) make the large-scale

sequencing possible. These techniques have been incorporated into the genome

sequencing and successfully sequenced the first bacterial genome (Fleischmann et al.

1995). Although the cost and time decreased dramatically, the principle technique

37

involved in genome sequence was still the Sanger method (Sanger et al. 1973). Recently,

a new pyrosequencing technology commercialized by Roche was developed to sequence

and generate many short pieces of DNA that could be assembled by the so-called 454

method. The first generation 454 method can read around 100 bp per read, and the read

number was increased to ~500 bases recently. 454 methods could finish sequencing a

bacterial genome within a few hours (Margulies et al. 2005) and at a relatively low cost.

The results are many contigs with different size fragment and there are some gaps left

behind. The quality of results mainly depends on the properties of the genome and the

number of runs performed. Of course, from the same template sample, more runs can

assure a better quality with less number of contigs at a higher price. The fosmid library

could be constructed and used to order, direct and space the contigs to form a bigger

“assembly” (Goldberg et al. 2006).

2.4.2 Microarray

DNA microarray or gene chips have been widely used for the functional genomics research. The availability of genome sequence is a prerequisite for the microarray construction. Therefore, most functional genomics studies are performed within eukaryotic species, especially for human medical research. Only several studies have been conducted with bacteria, such as B. subtilis, C. acetobutyricum, C. glutamicum, E. coli (Ikeda and Nakagawa 2003; Alsaker and Papoutsakis 2004; Gaertner et al. 2004).

However, for most of important bacteria in industry, the genomic sequence information is

38 limited. In such case, a shot-gun DNA microarray may be employed to investigate the functional genomics.

In the shotgun microarray approach, the whole genome is represented by random

DNA fragments of unknown sequence. These fragments are printed on a specially treated surface. The DNA fragments containing the target gene can be identified by the differential hybridization with dye-labeled cDNA from different conditions. Shotgun

DNA microarray have been constructed and used to study the metabolism of archaeon

Haloferax volvanii (Zaigler et al. 2003), gene function in environmental isolates of

Leptospirillum ferrooxidans (Parro and Moreno-Paz 2003), and phylogenetic lineages of

Listeria monocytogenes (Zhang et al. 2003).

39

2.5 References

Alam, S.;D. StevensR. Bajpai (1988). "Production of butyric acid by batch fermentation of cheese whey with Clostridium beijerinckii." J. Indus. Microb. 2(6): 359-364.

Alasfour, F. N. (1997). "Butanol-A single cylinder engine study: engine performance." Int. J. Hydrog. Energy 21: 21-30.

Alsaker, K.E. T. Papoutsakis (2004). Microarray-based analysis of the stress response in Clostridium acetobutylicum cultures. 227th ACS National Meeting. Anaheim, CA.

Atsumi, S.;A. F. Canna;M. R. Connora;C. R. Shena;K. M. Smitha;M. P. Brynildsena;K. J. Y. Choua;T. HanaiJ. C. Liao (2008). "Metabolic engineering of Escherichia coli for 1- butanol production." Metabo. Eng. 10(6): 305-311.

Bai, M. D.;S. S. ChengY. C. Chao (2004). "Effects of substrate components on hydrogen fermentation of multiple substrates." Water Sci. Tech. 50(8): 209-216.

Biebl, H. (2001). "Fermentation of glycerol by Clostridium pasteurianum-batch and continuous culture studies." J. Indus. Microb. & Biotech. 27(1): 18-26.

Bisson, L. F. (1988). "High affinity glucose transport in Saccharomyces cerevisiae is under general glucose repression control." J. Bacteriol 170: 4838-4835.

Bisson, L. F.;D. M. Coons;A. L. KruckebergD. A. Lewis (1993). "Yeast sugar transporters." Crit. Rev. Biochem. Mol. Biol. 28: 259-308.

Book, C.-W. F. The world fact book. CIA, https://www.cia.gov/cia/publications/factbook/rankorder/2174rank.html.

Boyaval, P.;C. CorreS. Terre (1987). "Continuous lactic acid fermentation with concentrated product recovery by ultrafiltration and electrodialysis." Biotech. Lett. 9(3): 207-212.

Boyles, D. (1984). Bioenergy technology thermodynamics and costs. New York, Wiley.

Boynton, Z.;G. BennettR. Rudolph (1996). "Cloning, sequencing, and expressing of genes encoding phosphotransacetylase and acetate kinase from C. acetobutylicum ATCC824." Appl. and Enviro. Microbiol. 62: 2758-2766.

40

Bro, C.;B. Regenberg;J. ForsterJ. Nielsen (2006). "In silico aided metabolic engineering of Saccharomyces cerevisiae for improved bioethanol production." Metabo. Eng. 8: 102- 111.

Bro, C.;B. RegenbergJ. Nielsen (2004). "Genome-wide transcriptional response of a Saccharomyces cerevisiae strain with an altered redox metabolism." Biotechnology and Bioengineering 85: 269-276.

Broder, J. D.;J. W. Barrier;K. P. LeeM. M. Bulls (1995). "Biofuels system economics." World Resour. Rev. 7(4): 560-569.

Brosseau, J. D.J. E. Zajic (1982). "Continuous microbial-production of hydrogen gas." Int. J. Hydrog. Energy 7(8): 623-628.

Buelter, T.;A. Hawkins;K. Kersh;P. Meinhold;M. PetersE. Subbian (2008). "Engineered microorganisms for producing n-butanol from glycerol." PCT Int. Appl.: 105.

Cammack, R.;M. FreyR. Robson (2001). Hydrogen as a fuel: Learning from future. London, Taylor & Francis.

Casey, G. P.W. E. Ingledew (1986). "Ethanol tolerance in yeasts." Crit. Rev. Microbiol 13: 219-280.

Cato, E.;W. GeorgeS. Finegold (1986). Genus clostridia prazmowski 1880. Bergey's manual of systematic bacteriology. Baltimore, William & Wilkins Co. 2: 1141-1200.

Chen, J. S.S. F. Hiu (1986). "Acetone butanol isopropanol production by Clostridium- Beijerinckii(Synonym: Clostridium Butylicum)." Biotechnol. Lett. 8: 371-376.

Cheng, S. S.;S. M. ChangS. T. Chen (2002a). "Effects of volatile fatty acids on a thermophilic anaerobic hydrogen fermentation process degrading peptone." Water Sci. Tech. 46: 209-214.

Cheng, S. S.;S. M. ChangS. T. Chen (2002b). Effects of volatile fatty acids on a thermophilic anaerobic hydrogen fermentation process degrading peptone. 2nd World Water Congress: Wastewater Treatment and Sludge Managment. 46: 209-214.

Chittibabu, G.;K. NathD. Das (2006). "Feasibility studies on the fermentative hydrogen production by recombinant Escherichia coli." Proc. Biochem. 41: 682-688.

Chu, C. F.;Y. Ebie;Y. InamoriH. N. Kong (2009). "Effect of hydraulic retention time on the hydrogen yield and population of Clostridium in hydrogen fermentation of glucose." J. of Enviromental Sciences (Beijing, China) 4: 424-428.

41

Classen, P.;J. Van Lier;C. Lopez;E. Van Niel;L. Sijtsma;A. Stams;S. De VriesR. Weusthuis (2000). "Utilization of biomass for the supply of energy carriers." Appl. Microbiol. Biotechnol. 52: 741-755.

Collet, C.;N. Adler;J. P. SchwitzguebelP. Peringer (2004). "Hydrogen production by Clostridium thermolacticum during continous fermentation of lactose." Int. J. Hydrog. Energy 29(14): 1479-1485.

Connor, M. R.J. C. Liao (2008). "Engineering of an Escherichia coli strain for the production of 3-methyl-1-butanol." Appl. Environ. Microbiol. 74(18): 5769-5775.

Cornillot, E.;R. V. Nair;E. T. PapoutsakisP. Soucaille (1997). "The genes for butanol and acetone formation in Clostrium acetobutylicum ATCC 824 reside on a large plasmid whose loss leads to degeneration of the strain." J. Bacteriol 179: 5442-5447.

Cornillot, E.P. Soucaille (1996). "Solvent forming genes in Clostridia." Nature 380: 489.

Crabbendam, P. M.;O. M. NeijsselD. W. Tempest (1985). "Metabolic and energetic aspects of the growth of Clostridium butyricum on glucose in chemostat culture." Archives of Microb. 142(4): 375-382.

Das, D.T. N. Veziroglu (2001). "Hydrogen production by biological processes: a survey of literature." Int J Energy Res 26(13-28).

Desai, R. P.;L. M. Harris;N. E. WelkerE. T. Papoutsakis (1999). "Metabolic flux analysis elucidates the importance of the acid-formation pathways in regulating solvent production by Clostridium acetobutylicum." Metabo. Eng. 1: 206-213.

Desai, R. P.E. T. Papoutsakis (1999). "Antisense RNA strategies for metabolic engineering of Clostridium acetobutylicum " Appl. Environ. Microbiol. 65: 936-945.

Dharmadi, Y.;A. MurarkaR. Gonzalez (2006). "Anaerobic fermentation of glycerol by Escherichia coli: a new platform for metabolic engineering." Biotech. Bioeng. 94(5): 821-829.

Dos Santos, M. M.;V. Raghevendran;P. Kotter;L. OlssonJ. Nielsen (2004). "Manipulation of malic enzyme in Saccharomyces cerevisiae for increasing NADPH production capacity aerobically in different cellular compartments." Metabo. Eng. 6: 353- 363.

Dziedzak, J. D. (1986). "Biotechnology and flavor development: An industrial research perspective " Food Technology 40: 8-20.

42

Elam, C.;C. Gregoire-Padro;G. Sandrock;A. Luzzi;P. LindbladE. Hagen (2003). "Realizing the hydrogen future: the International Energy Agency's efforts to advance hydrogen energy technologies." Int. J. Hydrogen Energy 28: 601-607.

Fabiano, B., Perego, P., ( 2002). " Thermodynamic study and optimization of hydrogen production by Enterobacter aerogenes." Int. J. Hydrog. Energy 27,: 149-156.

Fabiano, B., Perego, P., ( 2002..). " Thermodynamic study and optimization of hydrogen production by Enterobacter aerogenes." Int. J. Hydrog. Energy 27,: 149-156.

Fargione, J.;J. Hill;D. Tilman;S. PolaskyP. Hawthorne (2008). "Land cleaning and the biofuel carbon debt." Science 329(5867): 1235-1238.

Fleischmann, R. D.;M. D. Adams;O. White;R. A. Clayton;E. F. Kirkness;A. R. Kerlavage;C. J. Bult;J. F. Tomb;B. A. Dougherty;J. M. Merrick;K. Mckenney;G. Sutton;W. FitzHugh;C. Fields;J. D. Gocyne;J. Scott;R. Shirley;L. I. Liu;A. Glodek;J. M. Kelley;J. F. Weidman;C. A. Phillips;T. Spriggs;E. Hedblom;M. D. Cotton;T. R. Utterback;M. C. Hanna;D. T. Nguyen;D. M. Saudek;R. C. Brandon;L. D. Fine;J. L. Fritchman;J. L. Fuhrmann;N. S. M. Geoghagen;C. L. Gnehm;L. A. McDonald;K. V. Small;C. M. Fraser;H. O. SmithJ. C. Venter (1995). "Wholegenome random sequencing and assembly of haemophilus." Science 5223(269): 507-512.

Fontaine, L.;I. Meynial-Salles;L. Girbal;X. H. Yang;C. CrouxP. Soucaille (2002). "Molecular characterization and transcriptional analysis of adhE2, the gene encoding the NADH-dependent aldehyde/alcohol dehydrogenase responsible for butanol production in alcohologenic cultures of Clostridium acetobutylicum ATCC 824." J. Bacteriol 184(3): 821-830.

Frey, M. (2002). "Hydrogenases: Hydrogen-activationg enzymes." Chem. Bio. Chem. 3(2-3): 153-160.

Fritsch, M.;W. HartmeierJ. S. Chang (2008). "Enhancing hydrogen production of Clostridium butyricum using a cloumn reactor with square-structured ceramic fittings." Int. J. Hydrog. Energy 33(22): 6549-6557.

Fuhrmann, G.B. Volker (1992). "Regulation of glucose transport in Saccharomyces cerevisiae." J. Bacteriol 27: 1-15.

Gaertner, A. L.;H. Antelmann;R. Sapolsky;E. Ferrari;G. Chotani;B. MillerM. Hecher (2004). Qantitative proteomic and transcriptomic profiling during fermentation of pleiotropic Bacillus subtilis mutants. 227th ACS National Meeting. Anaheim, CA.

Goldberg, S.;J. Johnson;D. Busam;T. Feldblyum;S. Ferriera;R. Friedman;A. Halpern;H. Khouri;S. Kravitz;F. Lauro;K. Li;Y. Rogers;R. Strausberg;G. Sutton;L. Tallon;T.

43

Thomas;E. Venter;M. FrazierJ. Venter (2006). "A sanger/pyrosequencing hybrid approach for the generation of high-quality draft asseblies of marine microbial genomes." P.N.A.S. 103(30): 11240-11245.

Gorwa, M. F.;C. CrouxP. Soucaille (1996). "Molecular characterization and transcriptional analysis of the putative hydrogenase gene of Clostridium acetobutylicum ATCC 824." J. Bacteriol 178(9): 2668-2675.

Gray, C. T.H. Gest (1975). "Biological formation of molecular hydrogen." Science 148: 186-192.

Guettler, M. V.;M. K. JainB. K. Soni (1996). Process for making succinic acid, microorganisms for use in the process and methods of obtaining the microorganisms. US Patent 5504004. M. B. Institute.

Hallenbeck, P. C. (2005). "Fundamentals of the fermentative production of hydrogen." Water Science & Technology 52: 21-29.

Hallenbeck, P. C.J. R. Benemann (2002). "Biological hydrogen production: fundamentals and limiting processes." Int. J. Hydrogen Energy 27: 1185-1193.

Hanai, T.;S. AtsumiJ. C. Liao (2007). "Engineered synthetic pathway for isopropanol production in Escherichia coli." Appl. Environ. Microbiol. 73: 7814-7818.

Harris, L. M.;R. P. Desai;N. E. WelkerE. T. Papoutsakis (2000). "Characterization of recombinant strains of the Clostridium acetobutylicum butyrate kinase inactivation mutant: need for new phenomenological models for solventogenesis and butanol inhibition?" Biotech. Bioeng. 67: 1-11.

Hartimanis, M. G.S. Gatenbeck (1984). "Intermediary metabolism in clostridium acetobutylicum: levels of enzymes envolved in the fermentation of acetate and butyrate." Appl. Environ 47: 1277-1283.

Haug, M. (2004). Biofuels for transport; an international perspective. Paris, France, International Energy Agency.

Hawkes, F. R.;R. Dindale;D. L. HawkesI. Hussy (2002). "Sustainable fermentative biohydrogen: challenges for process optimization." Int J Energy Res 27: 1339-1347.

Hsu, T. A.;M. R. LadischG. T. Tsao (1980). "Alcohol from cellulose." Chem. Technol. 10(5): 315-319.

44

Huang, W. C.;D. E. RameyS. T. Yang (2004). "Continuous production of butanol by Clostridium acetobutylicum immobilized in a fibrous bed bioractor " Appl. Biochem. Biotechnol. 115: 887-898.

Huang, Y. L.;K. Mann;J. M. NovakS. T. Yang (1998). "Acetic acid production from fructose by Clostridium formicoaceticum immobilized in a fibrous-bed bioreactor." Biotech. Prog. 14: 800-806.

Huang, Y. L., Wu, Z.T., Zhang, L.K., Cheung, C.M., Yang, S.T. (2002). "Production of carboxylic acids from hydrolyzed corn meal by immobilized cell fermentation in fibrous- bed bioreactor." Bioresour Technol 82(1): 51-59.

Huang, Y. L.S. T. Yang (1998). "Acetate production from whey lactose using co- inmmobilized cells of homolactic and homoacetic bacteria in a fibrous-bed bioreactor." Biotech. Bioeng. 60: 499-507.

Huber, G. W.;S. IborraA. Corma (2006). "Synthesis of transportation fuels from biomass: chemistry, catalysis, and engineering." Chem. Rev. 106: 4044-4098.

Ikeda, M.S. Nakagawa (2003). "The corynebacterium glutamicum genome: features and impacts on biotechnological processes." Appl. Microbiol. Biotechnol. 62: 99-109.

Ingram, L. O. (1976). "Adaptation of membrane lipids to alcohols." J. Bacterial 125: 670- 678.

Ito, T.;Y. Nakashimada;K. Senba;T. MatsuiN. Nishio (2005). "Hydrogen and ethanol production from glecerol-containing wastes discharged after biodiesel manufacturing process." J. Biosci. Bioeng. 100(3): 260-265.

Ivey, D. M.L. G. Ljungdahl (1986). "Purification and characteristics of the F1-ATPase from Clostridium thermoaceticum." J. Bacterial 165: 252-257.

Jiang, L.;J. F. Wang;S. Z. Liang;X. N. Wang;P. L. CenZ. N. Xu (2009). "Butyric acid fermentation in a fibrous bed bioreactor with immoblilized Clostridium tyrobutyricum from cane molasses." Bioresour Technol 100(13): 3403-3409.

Jo, J. H.;D. S. LeeJ. M. Park (2008). "The effects of pH on carbon material and energy balances in hydrogen-producing Clostridium tyrobutyricum JM1." Bioresour Technol 99(17): 8485-8491.

Johnson, J.B. Francis (1975). "Taxonomy of the clostridia: ribosomal ribonucleic acid homologies among species." J. Gen. Microbial. 88: 229-244.

45

Johnson, J. L.J. S. Chen (1995). "Taxonomic relationships among strains of Clostridium acetobutylicum and other phenotypically similar organisms." FEMS Microbio. Rev. 17: 233-240.

Jones, D. T.S. Keis (1995). "Origins and relationships of industrial solvent-producing clostridial stains." FEMS Microbio. Rev. 17: 223-232.

Jones, D. T.D. R. Woods (1986). "Acetone-butanol fermentation revisited." Microb. Rev. 50(4): 484-524.

Jones, S. F.;G. M. EvansK. P. Galvin (1999). "The cycle of bubble production from a gas cavity in a supersaturated solution." Adv. Colloid and Inerface Sc. 80(1): 51-84.

Jungermann, K.;G. Leimenstoll;E. RupprechrtR. K. Thauer (1971). "Demonstration of NADH-ferrodoxin reductase in two saccharolytic clostridia." Archives fuer Mikrobiologie 80(4): 370-372.

Kakuda, H.;K. Hosono;K. ShiroshiS. Ichihara (1994). "Identification and characterization of the ackA (acetate kinase A)-pta (phosphotransacetylase) operon and complementation analysis of acetate utilizaiton by an ackA-pta deletion mutant of Escherichia coli." J. Biochem. 116: 916-922.

Kalscheuer, R.;T. StoeltingA. Steinbuechel (2006). "Microdiesel: Escherichia coli engineered for fuel production." Microbiology (Reading, United Kingdom) 152(9): 2529- 2536.

Karube, I.;N. Urano;T. Yamada;H. HirochikaK. Sakaguchi (1983). "Cloning and expression of the hydrogenase gene from Clostridium butyricum in Escherichia coli." FEBS Lett. 158: 119-122.

Kataoka, N.;A. MiyaK. Kiriyama (1997). "Studies on hydrogen production by continuous culture system of hydrogen producing anaerobic bacteria." Water Science & Technology 36: 41-47.

Kayser, A.;J. Weber;V. HechtU. Rinas (2005). "Metabolic flux analysis of Escherichia coli in glucose-limited continuous culture. I. Growth-rate dependent metabolic efficiency at steady state." Microbiology 151: 693-706.

Keis, S.;C. F. Bennett;V. K. WardD. T. Jones (1995). "Taxonomy and phylogeny of industrial solvent-producing clostridia." int. J. Syst. Bacterial. 45: 693-705.

Keis, S.;R. ShaheenD. T. Jones (2001). "Emended descriptions of Clostridium acetobutylicum and Clostridium beijerinckii and descriptions of Clostridium

46

saccharoperbutylacetonicum sp. nov. and Clsotridium saccharobutylicum sp. nov." Int. J. Syst. Evo. Microb. 51: 2095-2103.

Keshav, A.;K. L. WasewarS. Chand (2009). "Extraction of acrylic, propionic, and butyric acid using Aliquat 336 in oleyl alcohol: equilibria and effect of temperature." Industrial & Engineering Chemistry Research 48(2): 888-893.

Killeffer, D. H. (1927). "Butanol and acetone from corn." J. Indus. Eng. Chem. 19: 46-50.

Klass, D. L. (1998). Biomass for renewable energy, fuels and chemicals. San Diego, CA, Academic Press.

Klijn, N.;F. Nieuwenhof;J. Hoolwerf;V. d. W. C.A. Weerkamp (1995). "Identification of Clostridium tyrobutyricum as the causative agent of late blowing in cheese by species- specific PCR amplification." Appl. and Enviro. Microbiol. 61(8): 2919-2924.

Kraemer, J. T.D. M. Bagley (2007). "Improving the yield from fermentative hydrogen production." Biotechnol. Lett. 29: 685-695.

Kroschwitz, J. I.M. E. Howe-Grant (1997). Sugar to thin films. Kirk-Othermer Encyclopedia of Chemical Technology. Fourth Edition. 23: 1118.

Kumar, N.;A. GhoshD. Das (2001). "Redirection of biochemical pathways for the enhancement of H2 production by Enterobacter cloacae." Biotechnol. Lett. 23: 537-541.

Kumar, P.;D. M. Barrett;M. J. DelwicheP. Strove (2009). "Methods for pretreatment of ligncellulosic biomass for effeient hydrolysis and biofuel production." Industrial & Engineering Chemistry Research 48(8): 3713-3729.

Ladygina, N.;E. G. DedyukhinaM. B. Vainshtein (2006). "A review on micorbial systhesis of hydrocarbons." Proc. Biochem. 41(5): 1001-1014.

Lagunas, R. (1993). "Sugar transport in Saccharomyces cerevisiae " FEMS Microbio. Rev. 104: 229-242.

Larger, S. T.;S. Long;J. D. Santangelo;D. T. JonesD. R. Woods (1985). "Immobilized Clostridium acetobutylicum P262 mutants for solvent production." Appl. Environ. Microbiol. 50: 477-481.

Latimer, M. T.J. G. Fery (1993). "Cloning, sequence analysis, and hyperexpression of the genes encoding phosphotransacetylase and acetate kinase from Methanosarcina thermophila." J. Bacterial 175: 6822-6829.

47

Lee, S. M.;M. O. Cho;C. H. Park;Y. C. Chung;J. H. Kim;B. I. SangY. Um (2008). "Continuous butanol production using suspended and immobilized Clostridium beijerinckii NCIMB 8052 with supplementary butyrate." Energy & Fuels 22(3459-3464).

Lee, Y. J. (2004). "Effects of pH on continuous hydrogen fermentation." Hangug Hwangyeong bogeon Haghoeji 30(2): 149-153.

Lee, Y. J.;T. MiyaharaT. Noike (2001). "Effect of iron concentration on hydrogen fermentation." Bioresour Technol 80(3): 227-231.

Lee, Y. J.;T. MiyaharaT. Noike (2002). "Effect of pH on microbial hydrogen fermentation." J. Chem. Technol. Biotechnol. 77(6): 694-698.

Lemke, H. (1963). "New catalyst cycle in the oxo synthesis." Genie Chimique 89: 118- 120.

Levin, D. B.;L. PittM. Love (2004). "Biohydrogen production:prospects and limitations to practical application." Int J Energy Res 29: 173-185.

Li, M. Z.S. J. Elledge (2007). "Harnessing homologous recombination in vitro to generate recombinant DNA via SLIC." Nature Methods 4(3): 251-256.

Lienhardt, J.;J. Schripsema;N. QureshiH. P. Laschek (2002). "Butanol production by Clostridium beijerinckii BA101 in an immobilized cell biofilm reactor." Appl. Biochem. Biotechnol. 98: 591-598.

Lin, S. K.;C. Y. Du;A. Koutinas;R. H. WangC. Webb (2008). "Substrate and product inhibition kinetics in succinic acid production by Actinobacillus succinogenes." Biochem. Eng. J. 41: 128-135.

Linden, J. C.A. Moreira (1982). "Anaerobic production of chemicals." Biological basis for new developments in biotechnology 25: 377-403.

Liu, X.;Y. ZhuS. T. Yang (2006). "Construction and characterization of ack deleted mutant of Clostridium tyrobutyricum for enhanced butyric acid and hydrogen production." Biotech. Prog. 22: 1265-1275.

Liu, X. G.S. T. Yang (2006). "Kinetics of butyric acid fermentation of glucose and xylose by Clostridium tyrobutyricum wild type and mutant." Proc. Biochem. 41(4): 801-808.

Liu, Y. P.;P. Zheng;Z. H. Sun;Y. Ni;J. J. DongL. L. Zhu (2008). "Economical succinic acid production from cane molasses by Actinobacillus succinogenes." Bioresour Technol 99: 1736-1742.

48

Maeda, T.;V. Sanchez-TorresT. K. Wood (2007). "Enhanced hydrogen production from glucose by a metabolically-engineered Escherichia coli." Appl. Environ. Microbiol. 77(4): 879-890.

Magnusson, L.;R. Islam;R. Sparling;D. LevinN. Cicek (2008). "Direct hydrogen production from cellulosic waste materials with a single-step dark fermentation process." Int. J. Hydrog. Energy 33(20): 5398-5403.

Mahyudin, A. R.;Y. Furutani;Y. Nakashimada;T. KakizonoNishio (1997). "Enhanced hydrogen production n altered mixed acid fermentation of glucose by Enterobacter aerogenes." J. Ferm. Bioeng. 83(4): 358-363.

Maiorov, D. M.;D. V. MushenkoM. P. Vysotskii (1971). "Production of alcohols, acids, and aldehydes based on petroleum." Chem. Technol. of Fuels 7: 495-502.

Margulies, M.;M. Egholm;W. E. Altman;S. Attiya;J. S. Bader;L. A. Bemben;J. Berka;M. S. Braverman;Y. J. Chen;Z. Chen;S. B. Dewell;L. Du;J. M. Fierro;X. V. Gomes;B. C. Godwin;W. He;S. Helgesen;C. H. Ho;G. P. Irzyk;S. C. Jando;M. L. Alenquer;T. P. Jarvie;K. B. Jirage;J. B. Kim;J. R. Knight;J. R. Lanza;J. H. Leamon;S. M. Lefkowitz;M. Lei;J. Li;K. L. Lohman;H. Lu;V. B. Makhijani;K. E. McDade;M. P. McKenna;E. W. Myers;E. Nickerson;J. R. Nobile;R. Plant;B. P. Puc;M. T. Ronan;G. T. Roth;G. J. Sarkis;J. F. Simons;J. W. Simpson;M. Srinivasan;K. R. Tartaro;A. Tomasz;K. A. Vogt;G. A. Volkmer;S. H. Wang;Y. Wang;M. P. Weiner;P. Yu;R. F. BegleyJ. M. Rothberg (2005). "Genome sequencing in microfabricated high-density picolitre reactors." Nature 437(7057): 376-380.

Matsuyama, H.;H. YamamotoE. Nakano (1989). "Cloning, expression, and nucleotide sequences of the Escherichia coli. K-12 ackA gene." J. Bacterial 171: 77-80.

Meyer, J.J. Gagnod (1991). "Primary structure of hydrogenase I from Clostridium pasteurianumt." Biochem. 30: 9697-9704.

Michalodimitrakis, K.M. Isalan (2009). "Engineering prokaryotic gene circuits." FEMS Microbio. Rev. 33(1): 27-37.

Michel-Savin, D.;R. MarchalJ. P. Vandecasteele (1990a). "Butyrate production in continuous culture of Clostridium tyrobutyricum: effect of end-product inhibition." Appl. Microbio. Biotech. 33(2): 127-131.

Michel-Savin, D.;R. MarchalJ. P. Vandecasteele (1990b). "Control of the selectivity of butyric acid production and improvement of fermentation performance with Clostridium tyrobutyricum." Appl. Microbio. Biotech. 32(4): 387-392.

49

Michel-Savin, D.;R. MarchalP. Vandecasteele (1990c). "Butyric fermentation; Metabolic behavior and production performance of Clostridium tyrobutyricum in a continuous culture with cell recycle." Appl. Microbiol. Biotechnol. 34(172-177).

Minteer, S. D. (2006). Alcoholic fuels: An overview. Boca Raton, CRC Press.

Mishra, J.;S. Khurana;N. Kumar;A. K. GhoshD. Das (2004). "Molecular cloning, characterization, and overexpression of a novel [Fe]-hydrogenase isolated from a high rate of hydrogen producing Enterobacter cloacae IIT-BT08." BioChem. Biophys. Res. Commun. 324: 679-685.

Mitchell, W. J. (1998). "Physiology of carbohydrate to solvent conversion by clostridia." Adv. Microb. Physiol. 39: 31-130.

Mosier, N. S.;C. Wyman;B. Dale;R. Elander;Y. Y. Lee;M. HoltzappleM. R. Ladisch (2005). "Features of promising technologies for pretreatment of lignocellulosic biomass." Bioresour Technol 96: 673-686.

Mullis, K.;F. Faloona;S. Scharf;R. Saiki;G. HornH. Erlich (1986). "Specific enzymatic amplification of DNA in vitro: the chain reaction." Cold Spring Harb. Symp. Quant Biol. 51(Pt.1): 263-273.

Nair, R. V.E. T. Papoutsakis (1994). "Expression of plasmid-encoded aad in clostridium M5 restores vigorous butanol production." Journal of Bacteriology 176: 5843.

Nath, K.D. Das (2004). "Improvement of fermentative hydrogen production: various approaches." Appl. Microbio. Biotech. 65: 520-529.

News, C. E. Production report. Chem. Eng. News, ACS. July 11, 2005 83: 67-76.

Noike, T.;H. Takabatake;O. MizunoM. Ohba (2002). "Inhibition of hydrogen fermentation of organic wastes by lactic acid bacteria." Int J Energy Res 27: 1367-1371.

Nolling, J.;G. BretonM. V. Omelchenko (2001). "Genome sequence and comparative analysis of the solvent-producing bacterium Clostridium acetobutylicum." J. Bacteriol 183: 4823-4838.

Ozkan, P.;B. Sariyar;F. O. Utkur;U. AkmanA. Hortacsu (2005). "Metabolic flux analysis of recombinant protein overproduction in Escherichia coli." Biochem. Eng. J. 22: 167- 195.

Parro, V.M. Moreno-Paz (2003). "Gene function analysis in environmental isolates: the nif regulon of the strict iron oxidazing bacterium Leptospirillum ferrooxidans." Proc. Natl. Acad. Sci. USA 100: 7883-7888.

50

Patel, G. B.B. J. Agnew (1988). "Growth and butyric acid production by Clostridium populeti." Achives of Microb. 150(3): 267-271.

Penfold, D. W.;C. F. ForsterL. E. Macaskie (2003). "Increased hydrogen production by Escherichia coli strain HD701 in comparison with the wild-type parent strain MC4100." Enzyme Microb. Technol. 33: 185-189.

Perez, M.;K. Luyten;R. Michel;C. RiouB. Blondin (2005). "Analysis of Saccharomyces cerevisiae hexose carrier expression during wine fermentation: both low- and high- affinity Hxt transporters are expressed." FEMS Yeast Res. 5: 351-361.

Perlack, R. D.;L. L. Wright;A. Turhollow;R. L. Graham;B. StokesD. C. Erbach (2005). Biomass as feedstock for a bioenergy and bioproducts industry: The technical feasibility of a billion-ton annual supply. Oak Ridge, TN, Oak Ridge National Laboratory. DOE/GO-102995-2135.

Pfleger, B. F.;D. J. Pitera;C. D. SmolkeJ. D. Keasling (2006). "Combinatorial engineering of intergenic regions in operons tunes expression of multiple genes." Nature Biotech. 24(8): 1027-1032.

Playne, M. J. (1985). "Determination of ethanol, volatile fatty acids, lactic and succinic acids in fermentation liquids by gas chromatography." J. Sci. of Food Agri. 36(8): 638- 644.

Pouillart, P. R. (1998). "Role of butyric acid and its derivatives in the treatment of colorectal cancer and homoglobinopathies." Life Sci. 63(20): 1739-1760.

Price, N. D.;J. L. ReedB. O. Palsson (2004). "Genome-scale models of microbial cells: evaluating the consequences of constraints." Nature Rev. Microb. 2(11): 886-897.

Pryde, E. H. (1978). "Carboxylic acids-survey." Kirk-Othermer Encycl. Chem. Technol. 3rd Ed. 4: 814-834.

Rachman, M. A.;Y. Furutani;Y. Nakashimada;T. KakizonoN. Nishio (1997). "Enhanced hydrogen production in mixed acid fermentation of glucose by Enterobacter aerogenes." J. Fermen. and Bioeng. 83: 358-363.

Ramey, D. E. (1998). "Continuous two stages, dual path anaerobic fermentation of butanol and other organic solvents using two different strains of bacteria." US patent 5753474.

51

Rephaeli, A.;R. ZhukA. Nudelman (2000). "Prodrugs of butyric acid from bench to bedside: synthetic design, mechanisms of action, and clinical applications." Drug Development Resear. 50(3/4): 379-391.

Riebeling, V.K. Jungermann (1976). "Properties and function of Clostridial membrane ATPase." Biochem. Biophys. Acta 430(3): 434-444.

Sanchez, A. M.;J. Andrews;I. Hussein;G. N. BennettK. Y. San (2006). "Effect of overexpression of a soluble pyridine nucleotide transhydrogenase (UdhA) on the production of poly(3-hydroxybutyrate) in Escherichia coli." Biotechnol. Prog. 22: 420- 425.

Sanger, F.;G. M. Air;B. G. Barrell;N. L. Brown;A. R. Coulson;C. A. Fiddes;C. A. Hutchison;P. M. SlocombeM. Smith (1977). "Nucleotide sequence of bacteriophage phi X174 DNA." Nature 265(5596): 687-695.

Sanger, F.;J. E. Donelson;A. R. Coulson;H. KosselD. Fischer (1973). "Use of DNA polymerase I primed by a synthetic oligonucleotide to determine a nucleotide sequence in phage f1 DNA." Proc. Natl. Acad. Sci. U.S.A. 70(4): 1209-1213.

Schoondermark-Stolk, S. A.;M. Jansen;J. H. Veurink;A. J. Verkleij;C. T. Verrips;G. J. Euverink;J. BoonstraL. Dijkhuizen (2006). "Rapid identification of target genes for 3- methyl-1-butanol production in Saccharomyces cerevisiae." Appl. Microbio. Biotech. 70: 237-246.

Sentheshanmuganathan, S. (1960). "Mechanism of the formation of higher alcohols from amino acids by Saccharomyces cerevisiae." Biochem. J. 74: 568-576.

Shetty, R. P.;D. EndyT. F. Knight (2008). "Engineering biobrick vectors from biobrick parts." J. Bio. Eng. 2(5): doi:10.1186/1754-1611-2-5.

Shinto, H.;Y. Tashiro;M. Yamashita;G. Kobayashi;T. Sekiguchi;T. Hanai;Y. Kuriya;M. OkamotoK. Sonomoto (2007). "Kinetic modeling and sensitivity analysis of acetone- butanol-ethanol production." J. Biotech. 131: 45-56.

Silva, E. M.S. T. Yang (1995). "Kinetics and stability of a fibrous-bed bioreactor for continuous production of lactic acid from unsupplemented acid whey." J. Biotechnol. 41: 59-70.

Stephanopoulos, G. N.;A. A. AristidouJ. Nielsen (1998). Metabolic Engineering: Principles and Methodologies. San Diego, USA, Academic Press.

52

Tangney, M.;A. Galinier;J. DeutscherW. J. Mitchell (2003). "Analysis of the elements of catabolite repression in Clostridium acetobutylicum ATCC 824." J. Microb. Biotech. 6: 6-11.

Thauer, R. K.;K. JungermannK. Decker (1977). "Energy conservation in chemotrophic anaerobic bacteria." Bioteriological Reviews 41(1): 80-100.

Tolvanen, K.;P. Koskinen;H. Raussi;A. Ylikoski;I. Hemmilae;V. SantalaM. Karp (2008). "Profiling the hydA gene and hydA gene transcript levels of Clostridium butyricum during continuous, mixed-culture hydrogen fermentation." Int. J. Hydrog. Energy 33(20): 5416-5421.

Tomas, C. A.;K. V. Alsaker;H. P. Bonarius;W. T. Henderiksen;H. Yang;J. A. Beamish;C. J. ParedesE. T. Papoutsakis (2003). "DNA array-based transcriptional analysis of asporogenous, nonsolventogenic Clostridium acetobutylicum strains SKO1 and M5." J. Bacteriol 185: 4539-4547.

Tummala, S. B.;S. G. JunneE. T. Papoutsakis (2003). "Antisense RNA downregulation of coenzyme A transferase combined with alcohol-aldehyde dehydrogenase overexpression leads to predominantly alcohologenic Clostridium acetobutylicum fermentations " J. Bacteriol 185: 3644-3653.

Turner, J.;G. Sverdrup;M. K. Mann;P. C. Maness;B. Kroposki;M. Ghirardi;R. J. EvansD. Blake (2008). "Renewable hydrogen production." Int J Energy Res 32(5): 379-407.

Urbance, S. E.;A. L. Pometto III;A. A. DiSpititoY. Denli (2004). "Evaluation of succinic acid continuous and repeat-batch biofilm fermentation by Actinobacillus succinogenes using plastic composite support bioreactors." Appl. Microbio. Biotech. 65: 664-670.

Valentine, R. C.R. S. Wolfe (1960). "Purification and role of phosphotransbutyrylase." J. Bio. Chem. 235: 1948-1952.

Van Groenestijn, J. W.;J. H. Hazewinkel;M. NienoordP. J. Bussmann (2002). "Energy aspects of biological hydrogen production in high rate bioreactors operated in the thermophilic temperature range." Int. J. Hydrog. Energy 27(11): 1141-1147.

Varma, A.B. O. Palsson (1994). "Metabolic flux balancing- basic concepts scientific and practical use." Bio-Technol. 12: 994-998.

Vignais, P. M.;B. BilloudJ. Meyer (2001). "Classification and phylogeny of hydrogenases." FEMS Microbio. Rev. 25(4): 455-501.

Vignais, P. M.A. Colbeau (2004). "Molecular biology of microbial hydrogenases." Curr. Issues in Mol. Bio. 6(2): 159-188.

53

Walsh, M. C.;H. P. Smits;M. ScholteK. Van Dam (1994). "Affinity of glucose transport in Saccharomyces cerevisiae is modulated during growth on glucose." J. Bacteriol 176: 953-958.

Walter, K.;R. Nair;J. CaryG. Bennett (1993). "Sequence and arrangement of two genes of the butyrate-synthesis pathway of Clostridium acetobutylicum ATCC824." Gene 134: 107-111.

Wang, F. H.;H. J. Qu;D. W. Zhang;P. TianT. W. Tan (2007). "Produciton of 1,3- propanediol from glycerol by recombinant E.coli using incompatible plasmids system." Mol. Biotech. 37(2): 112-119.

Wang, M. Y.;Y. L. Tsai;B. H. OlsonJ. S. Chang (2008). "Monitoring dark hydrogen fermentation performance of indigenous Clostridium butyricum by hydorgenase gene expression using RT-PCR and qPCR." Int. J. Hydrog. Energy 33(18): 4730-4738.

Weyer, E. R.L. F. Rettger (1927). "A comparative study of six different strains of the organism commonly concerned in large-scale production of butyl alcohol and acetone by the biological process." J. Bacteriol 14: 399-424.

Whang, L. M.;C. J. HsiaoS. S. Cheng (2006). "A dual-substrate steady-state model for biological hydrogen production in an anaerobic hydrogen fermentation process." Biotechnol. Bioeng. 95(3): 492-500.

Williams, E. A.;J. M. CoxheadJ. C. Mathers (2003). "Anti-cancer effects of butyrate: use of microarray technology to investigate mechanisms." Proc. Nutr. Soc. 62: 107-115.

Wisenborn, D. P.;F. B. RudolphE. T. Papoutsakis (1989). "Phosphotransbutyrylase from Clostridium acetobutylicum ATCC824 and its role in acidogenesis." Appl. and Enviro. Microbiol. 66: 317-322.

Woodward, J.;M. Mattingly;M. Danson;D. Hough;N. WardM. Adams (1996). "In vitro hydrogen production by glucose dehydrogenase and hydrogenase." Nat. Biotechnol. 14: 872-874.

Wu, Z. T.S. T. Yang (2003). "Extractive fermentation for butyric acid production from glucose by Clostridium tyrobutyricum." Biotech. Bioeng. 82: 93-102.

Yang, S. T. (1996). "Extractive fermentation using convoluted fibrous bed bioreactor." U.S. patent US 5563069

Yang, S. T.;H. Zhu;Y. LiG. Hong (1994). "Continuous propionate production from whey permeate using a novel fibrous bed bioreactor." Biotech. Bioeng. 43(11): 1124-1130.

54

Yat, S. C.;A. BergerD. R. Shonnard (2008). "Kinetic characterization for dilute sulfuric acid hydrolysis of timber varieties and switchgrass." Bioresour Technol 99(9): 3855-63.

Yoshida, A.;T. Nishimura;H. Kawaguchi;M. InuiH. Yukawa (2006). "Enhanced hydrogen production from glucose using Idh- and frd-inactivated Escherichia coli strains." Appl. Microbio. Biotech. 73: 67-72.

Yoshida, A.;T. Nishimura;H. Kawaguchi;M. InuiH. Yukawa (2007). "Efficient induction of formate hydrogen of aerobically grown Escherichia coli in a three-step biohydrogen production process." Appl. Microbio. Biotech. 74: 754-760.

Yoshimoto, H.;T. Fukushige;T. YonezawaH. Sone (2002). "Genetic and physiological analysis of branched-chain alcohols and isoamyl acetate production in Saccharomyces cerevisiae." Appl. Microbio. Biotech. 59: 501-508.

Young, M.S. Cole (1993). Bacillus subtilis and other gram-positive bacteria: biochemistry, physiology, and molecular genetics. Washington D.C., American Society for Microbiology.

Zaigler, A.;S. C. SchusterJ. Soppa (2003). "Construction and usage of a onefold-coverage shotgun DNA microaaray to characterize the metabolism of the archaeon Haloferax volcanii." Mol. Microbiol. 48: 1089-1105.

Zhang, C.;M. Zhang;J. Ju;J. Nietfeldt;J. Wise;P. M. Terry;M. Olson;S. D. Kachman;M. Wiedmann;M. SamadpourA. K. Benson (2003). "Genome diversification in phylogenetic lineages I and II of Listeria monocytogenes: Identification of segments unique to lineage II populations." J. Bacterial 185: 5573-5584.

Zhao, Y.;C. A. Tomas;F. B. Rudolph;E. T. PapoutsakisG. N. Bennett (2005). "Intracellular butyryl phosphate and acetyl phosphate concentrations in Clostridium acetobutylicum and their implications for solvent formation." Appl. Environ. Microbiol. 71: 530-537.

Zhu, Y.;X. LiuS. T. Yang (2004). "Construction and characterization of pta gene-deleted mutant of Clostridium tyrobutyricum for enhanced butyric acid fermentation." Biotech. Bioeng. 90: 154-166.

Zhu, Y.;Z. T. WuY. S.T. (2002). "Butyric acid production from acid hydrolysate of corn fiber by Clostridium tyrobutyricum in a fibrous-bed bioreactor." Proc. Biochem. 38(5): 657-666.

Zhu, Y.S. T. Yang (2003). "Adaptation of Clostridium tyrobutylicum for enhanced tolerance to butyric acid in a fibrous-bed bioreactor." Biotech. Prog. 19: 365-372.

55

Zigova, J.E. Sturdisk (2000). "Advances in biotechnological production of butyric acid." J. Indus. Microb. & Biotech. 24(3): 153-160.

Zigova, J.;E. Sturdisk;D. VandakS. Schlosser (1999). "Butyric acid production by Clostridium butyricum with integrated extraction and pertraction." Proc. Biochem. 34: 153-160.

56

Figure 2.1 The metabolic pathway in Clostridium acetobutylicum ATCC 824 (pfor: pyruvate: ferredoxin oxidoreductase; nfor: NADH: ferredoxin oxidoreductase; hyd: hydrogenase; pta: phosphotransacetylase; ack: acetate kinase; hbd: 3- hydroxybutyryryl CoA dehydrogenase; ptb: phosphotransbutyrylase; buk: butyrate kinase; ctf: CoA transferase; adc: acetoacetate decarboxylase; aad: aldehyde/alcohol dehydrogenase; bdh: butanol dehydrogenase; adh: alcohol dehydrogenase)

57

Figure 2.2 The metabolic pathway in Clostridium tyrobutyricum ATCC 25755 (pfor: pyruvate: ferredoxin oxidoreductase; nfor: NADH: ferredoxin oxidoreductase; hyd: hydrogenase; pta: phosphotransacetylase; ack: acetate kinase; hbd: 3- hydroxybutyryryl CoA dehydrogenase; ptb: phosphotransbutyrylase; buk: butyrate kinase)

58

CHAPTER 3

CONSTRUCTION AND CHARACTERAZATION OF ptb MUTANT OF

CLOSTRIDUM TYROBUTYRICUM

Summary

Clostridium tyrobutyricum is a rod-shape, gram-positive bacterium that produces

butyrate, acetate, hydrogen and carbon dioxide from various saccharides, such as glucose

and xylose, under strict anaerobic conditions. Phosphotransbutyrylase (PTB) is a key

enzyme in the butyric acid biosynthesis pathway. However, its role in affecting metabolic

flux distribution and production of various metabolites is not well understood. In this

work, a mutant with interrupted ptb gene, encoding phosphotransbutyrylase, was created

by homologous recombination with a non-replicative integrational plasmid carrying ptb

fragment disrupted by erythromycin resistant gene. Compared to the wild type, the activities of PTB and butyrate kinase (BK) in the mutant decreased 76% and 42%, respectively; meanwhile, phosphotransacetylase (PTA) and acetate kinase (AK) increased

7% and 29%. However, the disruption of ptb did not reduce butyric acid production from

glucose or xylose in batch fermentation. In addition, both acetic acid and hydrogen

production increased significantly. The mutant also displayed a higher specific growth

rate (0.20 h-1 vs. 0.15 h-1 for growth on glucose and 0.14 h-1 vs. 0.10 h-1 for growth on xylose) and higher tolerance to butyric acid (Kp = 13.5 vs. 9.0). Consequently,

59

fermentation in serum bottle gave a lower butyrate/acetate ratio and higher hydrogen yield by ptb mutant than wild type.

3.1 Introduction

C. tyrobutyricum is a member of clostridia family and was discovered as the causative agent for the late-blowing defect in cheese production. It produces butyric acid, acetic acid, CO2 and hydrogen from diverse carbon source including glucose and xylose

(Klijn et al. 1995). Butyric acid is a 4-carbon carboxylic acid with many applications in

food and perfume industries (Dziedzak 1986). Recent research suggested that butyrate

have the ability to inhibit the histone deacetylases and is being evaluated as anti-

neoplastic therapy (Pouillart 1998; Williams et al. 2003). The production of butyric acid

from biomass became an alternative to the current petroleum chemical route due to the

pollution and the depletion of the petroleum. Compared with other bacteria in butyric

acid fermentation, C. tyrobutyricum has many advantages, including simple medium for cell growth and high productivity and yield (Wu and Yang 2003). Furthermore, hydrogen

as gas by-product, is a clean, efficient and sustainable energy source with the highest

energy content per unit weight (143 GJ/ton) among all known fuels (Boyles 1984). In

addition, hydrogen does not contribute to any environmental problems since its oxidation

product is H2O only.

The suggested metabolic pathway in C. tyrobutyricum is shown in Figure 3.1. In this pathway, the substrate, hexose or pentose, is oxidized to acetyl-CoA, with the releasing of CO2 and hydrogen. Then, acetyl-CoA is either converted to acetyl phosphate

60 resulting in the excretion of acetate, or condensed and reduced to form butyryl-CoA resulting in the excretion of butyrate. The butyrate formation pathway is comprised of two consecutive enzymatic steps catalyzed by phosphotransbutyrylase (PTB) and butyrate kinase (BK), encoded by ptb and buk, respectively (Walter et al. 1993). Both enzymes have been purified and characterized from Clostridium acetobutylicum strain

ATCC 824 (Cary et al. 1988). PTB is an octamer of identical 31-kDa subunits consisting of 302 residues and BK was reported to be a dimer of identical 30-kDa subunits consisting of 356 residues. ptb and buk are located in the same operon with ptb proceeding buk 44 bp with two independent RBS (Hartimanis and Gatenbeck 1984;

Wisenborn et al. 1989). They played critical roles in butyric acid biosynthesis in C. acetobutylicum (Green et al. 1996; Green and Bennet 1997). In order to increase butyric acid production by C. tyrobutyricum, mutants with inactivated phosphotransacetylase

(PTA) and acetate kinase (AK), which catalyze the formation of acetate and also generate

ATP for cell growth (Green et al. 1996), were developed and they showed increased butyrate and decreased acetate production in batch fermentations (Zhu et al. 2004; Liu et al. 2006; Liu and Yang 2006). However, no work has been done on ptb and buk in C. tyrobutyricum.

The goal of this work was to understand the effects of ptb-disruption on butyric acid fermentation by C. tyrobutyricum. We were particularly interested in increasing hydrogen production in butyric acid fermentation by using ptb-disrupted C. tyrobutyricum. It was hypothesized that blocking the butyrate production pathway can redirect the carbon flux to acetate production and hence shift the reducing equivalent to

61

hydrogen production. The effects of ptb-disruption on butyric acid, acetic acid and

hydrogen production from glucose fermentation in serum bottles were studied and the

results are reported here. The potential of using ptb-disrupted mutant for butyrate and

hydrogen production are also discussed in this chapter.

3.2 Materials and Methods

3.2.1 Cultures and media

The bacterial strains and plasmids used in this study are listed in Table 3.1. Unless

otherwise noted, C. tyrobutyricum was maintained in RCM (Reinforced Clostridium

Medium, Difco) in anaerobic conditions at 4°C. The fermentation medium was CGM

(Clostridium growth medium) prepared as previously described by Huang et al. (Huang

et al. 1998) and supplemented with glucose or xylose as the carbon source. E. coli was

grown aerobically at 37°C in Luria-Bertani medium (LB) supplemented with 50 µg/ml

ampicillin.

3.2.2 Cloning of ptb-disrupted mutant

The chromosomal DNA of wild type of ATCC 25755 was prepared using the

QIAGEN genomic DNA kit. The amplification of the partial ptb sequence from C.

tyrobutyricum chromosome DNA and the construction of the non-replicative plasmid

used in the transformation to obtain ptb-disrupted mutant followed similar procedures

described by Zhu et al. (Zhu et al. 2004).

62

3.2.2.1 PCR amplification for ptb fragment.

The whole genomes of several clostridia species, including C. acetobutylicum, C. beijerinckii, C. tetani, and C. perferingens, have been sequenced and annotated (Nolling

et al. 2001; Shimizu et al. 2002; Bruggemann et al. 2003; Copeland et al. 2009). The

amino acid sequence of PTB is highly conserved among those species (Figure 3.3).

Degenerated primers were designed based on the highly conserved regions and the codon

usage preference for C. tyrobutyricum (http://www.kazusa.or.jp/codon). The sequences of

the degenerate primers were as follows:

Forward primer: 5’- AATA(T)A(G)TAAAC(T)GAA(G)CCTAACG

Reverse primer: 5’-A(G)CTA(G)TCA(T)GCTCTA(T)GAAGTTA

PCR amplification of the partial ptb gene from the wild type C. tyrobutyricum chromosomal DNA was performed with optimized PCR buffer containing 2.5 mM

MgCl2, 0.5 mM dNTP, and 0.3 mM primers (each). The optimized program was as

follows: 94°C for 3 min; 94°C for 50 s, 48°C for 50 s, 72°C for 1 min, 40 cycles,

followed by 72°C for 10 min to allow the addition of deoxyadenosine (A) to the 3’ ends

of PCR products. A 636 bp fragment was obtained and cloned into the pCR®4-TOPO vector (Invitrogen) to form plasmid pCR4ptb for sequencing and further manipulation.

3.2.2.2 Nested-PCR for the confirmation of C. tyrobutyricum chromosome

In order to confirm the chromosome template was from C. tyrobutyricum and no contamination from C. acetobutylicum, the 16S rRNA identification was performed by nested-PCR (Klijn et al. 1995). In the first step of nested PCR, part of 16S rRNA gene

63 was amplified by the Clostridium-specific primers at T = 55°C. The forward primer used was 5’-GCGGCGTGCCTAATACATGC-3’ and the reverse primer used was 5’-

GGGTTGCGCTCGTTGCGGGGA-3’. The second amplification was performed using the diluted product of the first step as template with species-specific reverse primers

(GGACTTCATCCATTACGGACTAAC for C. acetobutylicum ATCC824 and

CGCCTATCTCTAGGTTATTCAG for C. tyrobutyricum) and a common forward primer

(GGAATCTTCCACAATGGGCG) in V6 region. The PCR result was examined on agarose gel and stained with EB.

3.2.2.3 Construction of non-replicative plasmid pPtbEm

The non-replicative plasmid pPtbEm was constructed by inserting a selection marker into the middle of the ptb fragment (Figure 3.4). Erythromycin has been successfully used as a selection marker in engineering C. acetobutylicum (Green and

Bennet 1997) and C. tyrobutyricum (Zhu et al. 2004; Liu et al. 2006). The 1.6 kb erythromycin resistant gene cassette (Emr) was obtained from plasmid pDG647 (Bacillus

Genetic Stock Center). The pDG647 was digested by BamHI (Guerout-Fleury et al.

1995) and the 1.6 kb Emr fragment was purified by the gel extraction kit (Qiagen). The pCR4ptb was digested by bglII, dephosphorylated by CIAP and purified by Phenol:

Chloroform extraction. Emr fragment was ligated with the linearized pCR4ptb and the non-replicative plasmid pPtbEm was constructed. The pPtbEm was then transformed into

E. coli DH5α and cells were cultured on LB plate supplemented with 250 μg/ml erythromycin and 100 μg/ml ampicillin. The colonies on the plate were selected and

64

cultured in LB with 50 μg/ml ampicillin, and then cells were harvested to extract the plasmid using QIAGEN plasmid MINIprep kit.

3.2.2.4 Transformation

Electroporation was used to introduce plasmid into C. tyrobutyricum following

the procedures previously described by Zhu et al. (Zhu et al. 2004). All manipulations

were operated in an anaerobic chamber (Thermo Fisher Scientific model 1025). Cells were grown in 50 ml of RCM supplemented with 40 mM DL-threonine and harvested at

OD600=0.6 (Zhu et al. 2004; Jiraskova et al. 2005), then washed with 30 ml of ice-cold

electroporation SMP buffer twice and resuspended in 1 ml of SMP buffer in 0.4 cm

electroporation cuvette (Bio-Rad). The SMP buffer (pH 7.4) consisted of 270 mM

sucrose, 1 mM MgCl2, and 7 mM sodium phosphate. The MgCl2 solution was prepared and autoclaved separately before adding to the sucrose and phosphate solution to prevent precipitation. Two micrograms of ice-cold pPtbEm plasmid were added and the mixture was incubated on ice for 5 min. The transformation parameters were set as follows: 2.5 kV, 600 Ω, 25 AF. The transformed cells were transferred into pre-warmed RCM, incubated at 37°C for 2 hours, and then plated on RCM plates containing 40 µg/ml erythromycin. Plates were incubated in an anaerobic incubator at 37°C for 3–5 days to develop the mutant colonies.

3.2.2.5 PCR characterization of ptb mutant

65

PCR was used to confirm the insertion of Emr in the ptb gene in the mutant. The chromosome DNA of mutant was extracted using the QIAGEN genomic DNA kit. The ptb primers and erythromycin resistant gene primers were designed to detect the length of the ptb fragment and the existence of erythromycin resistant gene in the chromosome of the mutant. The ptb primers were designed based on the sequence of the ptb fragment and had the following sequences: 5’-ACAACTGGTGCAGAAGTTCC (ptb forward) and 5’-

GTAGAGCTTGTATCAACTGG (ptb reverse). The erythromycin resistant gene primers were designed using the Invitrogen Vector NTI software and had the following sequences: 5’-AAGAAGATATGATAGTTTATGGCGG (Emr forward) and 5’-

GCACAGTTCATTATCAACCAAACAA (Emr reverse).

3.2.3 Enzyme activity assays

Cells from 50 ml of wild type or ptb mutant culture were harvested by centrifugation and the cell pellets were stored at -80°C until use. Cells were washed with

25 ml of tris-HCl buffer (pH 7.4, 0.1 M) twice and resuspended in 5 ml buffer. The cell suspension was sonicated on ice using a sonic dismembrator (Fisher scientific model 100) at level 2 for 10 seconds with a 30 seconds interval, repeated 5 times. The cell debris was removed by centrifugation at 16,000 × g for 15 min at 4°C. The total protein concentration was measured by Biorad protein assay following the standard procedure for microtiter plates (Biorad).

The activities of AK and BK in cell extracts were measured by monitoring the formation of acyl phosphate from acetate and butyrate, respectively (Rose 1955).

66

Acylphosphate can react with hydroxylamine to give hydroxamic acid, which forms a

complex with ferric ion giving yellow color with absorbance at 540 nm (Lipmann and

Tuttle 1945; Cary et al. 1988). Enzyme activity was calculated based on the extinction

coefficient of 0.169 mM-1cm-1 (Cary et al. 1988). One unit of AK or BK enzyme activity is defined as the amount of enzyme producing 1 μmol of hydroxamic acid per minute.

PTA and PTB activities were measured by detecting the release of CoA from acetyl-CoA and butyryl-CoA, respectively (Andersch et al. 1983). The sulfyl group in CoA was

quantified by the absorbance at 405 nm in the presence of Ellman’s reagent, 5,5'-dithio-

bis(2-nitrobenzoic acid) (Ellman 1959). An extinction coefficient of 13.6 mM-1cm-1 was used for activity calculation. One activity unit of PTA or PTB is defined as the amount of enzyme converting 1 μmol of acyl-CoA or butyryl-CoA per minute under the reaction conditions. The specific activity of enzyme is defined as the unit of activity per mg of protein.

3.2.4 Butyrate tolerance

Butyrate tolerance was evaluated by investigating the inhibition effect of butyrate on cell growth in serum bottles containing media with various amounts of sodium butyrate (0, 5, 10, 15 g/l). Each serum bottle containing 100 ml of RCM was inoculated with 2 ml of freshly prepared C. tyrobutyricum culture in the exponential phase (OD600 =

1.0) and incubated at 37°C. Samples were taken at regular intervals to measure the OD600

with a spectrophotometer (Sequoia-Turner, model 340). The specific growth rates at various initial butyrate concentrations were determined from the growth curves.

67

3.2.5 Fermentation kinetics

The effect of ptb-disruption on fermentation profile was evaluated in serum

bottles. Each bottle containing 50 ml of CGM with 10 g/l glucose was inoculated with 1

ml of fresh culture and incubated at 37°C. Samples were taken at regular time intervals

and the final concentration of acetate and butyrate were analyzed with HPLC.

Butyrate/acetate ratio was calculated by the mass ratio of butyrate and acetate to evaluate

the effects of ptb-disruption on the metabolic flux distribution.

3.2.6 Analytical methods

Cell density was calculated based on the optical density of the cell suspension at

600 nm (OD600) measured with a spectrophotometer (Sequoia-Turner, model 340). 0.68 g/l of cell dry weight corresponded to one unit of OD600. The organic contents, including glucose, xylose, butyrate, and acetate, in fermentation broth samples were analyzed with a high performance liquid chromatography (HPLC) with an organic acid analysis column

(Aminex HPX-87H, Biorad) following the method described in appendix C (Zhu and

Yang 2003).

3.3 Results

3.3.1 Cloning of ptb mutant

An estimated 650 bp of DNA fragment was amplified using the degenerate primers and the sequencing result showed a 636 bp fragment encoding a 212 amino acids

68

peptide. The partial ptb sequence isolated from C. tyrobutyricum genome had the

identical DNA sequence to the ptb sequence previously reported for C. acetobutylicum

ATCC 824 (gene ID 1119259, E.C. 2.1.3.19) (Walter et al. 1993).

Genus specific and species specific primers have been used to confirm that the

chromosome DNA is from C. tyrobutyricum. The Clostridium genus specific primers gave a band of 614 kb. The nested PCR using the previous round PCR product as template and use C. tyrobutyricum and C. acetobutylicum specific primers as primer separately, only C. tyrobutyricum specific primers gave expected band of 614 bp (Figure

3.3).

The protoplast of C. tyrobutyricum was prepared and incubated with pPtbEm plasmid to check the existence of restriction system. The result showed that the pPtbEm were not digested by the protoplast of C. tyrobutyricum (data not shown). Due to the difficulty of transformation of Clostridium, 40 µg/ml of DL-threonine were added into the medium to weaken the cell wall formation. Five colonies were isolated under the 40

μg/ml erythromycin selective pressure. The transformation efficiency was 2.5 /µg DNA, which was similar to the previous C. tyrobutyricum transformation result (Zhu et al.

2004). As a negative control, C. tyrobutyricum was also transformed with pDG647 plasmid and no colony was obtained. Since both pDG647 and pPtbEm contained the same replicon colE, the results confirmed that the pPtbEm cannot replicate in C. tyrobutyricum without integration into the chromosome. Also, the existence of ptb fragment in the non-replicative plasmid was necessary for the homologous recombination.

69

The homologous recombination of the pPtbEm with chromosome of C. tyrobutyricum might happen in two ways (Figure 3.5). One possibility is that the whole plasmid was inserted into the chromosome which would add 6.2 kb to the ptb fragment

(Zhu et al. 2004); while the other possibility was that only the Emr-inserted ptb fragment

exchanged with the ptb gene in the chromosome and this would cause the ptb fragment

1.6 kb longer. Both of the recombination could result in the inactivation of ptb gene

because a DNA fragment was inserted into the ptb gene. To confirm the insertion of the

foreign DNA fragment in the ptb gene, ptb primers were used to identify the lengths of

ptb fragment in the mutant and wild type. As expected, the wild type gave a 636 bp fragment while the mutant gave a 2.2 kb fragment, which is the length of the original ptb

plus the erythromycin resistant gene cassette (Figure 3.6A). A pair of Emr primers was

used to further confirm that the elongation of the ptb fragment was caused by the erythromycin resistant gene insertion (Figure 3.6B). The template was the PCR product of the mutant chromosome amplified by the ptb primers. It showed a single band with the same size as the positive control Emr-containing plasmid pDG647. As expected, there is

no erythromycin resistant gene found in wild type chromosome. These PCR results confirmed that the ptb gene in the chromosome of the mutant has been disrupted by the insertion of Emr.

3.3.2 Effects of ptb-disruption on acid-forming enzyme activities

The effects of ptb-disruption on the activities of acid-forming enzymes (PTB, BK,

PTA, and AK) in the mutant were investigated and the results are summarized in Figure

70

3.7. All enzyme activities are shown in the relative scale using wild type as 100%. PTB activity was reduced dramatically down to ~24% in the ptb-disrupted mutant. However, the ptb-disruption did not turn off the PTB activity completely suggesting the existence of isozyme such as PTA. Meanwhile, the BK activity also decreased ~42%, suggesting that buk is located at the downstream of ptb in the same operon (Walter et al. 1993). In contrast, AK activity increased ~29% in the mutant, but PTA activity was not significantly affected (p = 0.23).

3.3.3 Effect on fermentation profile

Three out of five clones obtained from the transformation grew well in the RCM containing 40 µg/ml erythromycin and all showed similar product profiles with a significantly (p = 0.007) reduced B/A ratio that was about 25.6% lower than that from the wild type fermentation (Figure 3.8). This result confirmed that the ptb-disruption caused a shift in the carbon flux toward acetate production. However, butyrate was still produced at a significant level, suggesting the existence of other butyrate biosynthesis pathway in this bacterium.

3.3.4 Butyrate tolerance

The butyrate tolerance was evaluated by investigating the effect of butyric acid on the specific growth rate. Cells were grown as suspension culture at various butyrate concentrations (0, 5, 10, 15 g/l). The relative specific growth rates with the rate without butyrate being 100% were plotted against the butyrate concentration (Figure 3.9). As

71

shown in Figure 9, butyrate strongly inhibited the wild type but not as strongly to the ptb mutant. At the highest butyrate concentration of 15 g/l, the ptb mutant displayed 53% of its maximum growth rate compared to 43% retained in the wild type. The growth inhibition by butyric acid followed the non-competitive product inhibition kinetics as follows (Liu et al. 2006):

μ K μ = max P P + PK

-1 -1 where µ is the specific growth rate (h ), µmax is the maximum specific growth rate (h ),

KP is the inhibition rate constant (g/l), and P is the butyric acid concentration (g/l). The

ptb mutant had a higher inhibition rate constant (Kp = 13.5 g/l vs. Kp = 9.0 g/l for the

wild type), suggesting that the ptb mutant has a higher tolerance to butyrate inhibition.

The higher growth rate and butyrate tolerance can be attributed to more ATP produced by

the ptb mutant, which could help H+-ATPase pumps more proton out of cells to maintain

a suitable pH in cells (O'Sullivan and Condon 1999; Cotter and Hill 2003).

3.4 Discussion

The Clostridium is a heterogeneous genus of gram-positive, obligate anaerobic, rod-shaped bacteria which can form endospore under extreme conditions. Clostridium attracts intense interest due to its wide applications in medical research and the fermentation industry. To date 6 clostridia species, C. perferingens (Shimizu et al. 2002),

C. difficile (Sebaihia et al. 2006), C. botulinum (Sebaihia et al. 2007), C. tetani

(Bruggemann et al. 2003), C. acetobutylicum (Nolling et al. 2001), and C. beijerinckii

72

(Copeland et al. 2009) have been fully sequenced for their whole genomes. However,

except for C. acetobutylicum, very little has been done in metabolic engineering of

clostridia for fuels and chemicals production. C. tyrobutyricum was originally discovered

as the causative agent for the late-blowing defect in cheese (Klijn et al. 1995). It produces

butyric acid, acetic acid, carbon dioxide, and hydrogen under anaerobic conditions.

Because of its many advantages in fermentation, its potential for industrial production of butyric acid and hydrogen from biomass has attracted intensive research interest (Michel-

Savin et al. 1990a; Michel-Savin et al. 1990b; Michel-Savin et al. 1990c; Zigova et al.

1999; Zhu et al. 2002; Wu and Yang 2003; Zhu and Yang 2003; Liu et al. 2006; Jo et al.

2008; Jiang et al. 2009). To improve butyric acid production, metabolically engineered C.

tyrobutyricum mutants with inactivated acetate formation pathway were obtained with

integrational mutagenesis that selectively inactivated ack and pta genes from the host

chromosome (Zhu et al. 2004; Liu et al. 2006). Fermentation kinetics with the mutants

showed that pta and ack inactivation increased butyrate production, reduced cell growth,

and decreased acetate production.

In this work, we were interested in cloning C. tyrobutyricum ptb gene for its

important role in the butyrate synthesis pathway. This pathway is important in the energy

metabolism of the organism, as ATP is produced during the conversion of butyryl-CoA to

butyrate (Valentine and Wolfe 1960). The partial ptb sequence of C. tyrobutyricum

obtained by PCR with degenerated primers has an identical DNA sequence to the one

found in C. acetobutylicum ATCC 824, which also has the same butyrate synthesis

pathway and produces butyric acid during its acidogenesis phase (Jones and Woods

73

1986). Previous studies showed that the partial DNA sequences of pta and ack genes from C. tyrobutyricum had high similarity, but not identical, to C. acetobutylicum (Zhu et al. 2004; Liu et al. 2006). In order to exclude the possibility of contamination of the chromosome with C. acetobutylicum in our cloning experiment, the chromosome’s C. tyrobutyricum origin was checked by amplifying its 16S rRNA gene following the method previously reported (Klijn et al. 1995). Briefly, clostridia-specific primers P1

(GCGGCGTGCCTAATACATGC) and P2 (GGGTTGCGCTCGTTGCGGGA) were used and the results showed its Clostridium origin. The PCR product was then used as the template for the next step of nested PCR with clostridia universal primer P5

(GGAATCTTCCACAATGGGCG) and C. tyrobutyricum specific primer Pty

(CGCCTATCTCTAGGTTATTCAG), and the specific band of 660 bp was obtained, confirming the C. tyrobutyricum origin of the chromosome. As a control, C. acetobutylicum specific primer Pac (GGACTTCATCCATTACGGACTAAC) did not give the amplification product.

Compared to the wild type, the ptb mutant’s PTB activity was reduced by 76% and BK activity decreased 42%. Meanwhile, its AK activity increased 29%. However, butyrate production was not reduced in the fermentation with the ptb mutant although acetate production did increase significantly. A mutant of C. acetobutylicum ATCC 824 with disrupted buk gene showed fivefold lower BK activity, twofold higher PTA activity and threefold higher AK activity (Green et al. 1996). Consequently, fermentation with this mutant showed a 43% decrease in butyrate and 32% increase in acetate production.

Meanwhile, the complementation of buk-inactivation strain with plasmid containing the

74

buk operon restored the butyrate kinase activity and butyrate production (Green and

Bennet 1997). These experiments showed that buk played an important role in butyrate

biosynthesis in C. acetobutylicum; however, its deletion could not eliminate butyrate

production, suggesting the existence of a new pathway other than PTB-BK route also

leading to butyrate biosynthesis in these clostridia species. Previous studies with ack and pta deleted mutants also suggested the existence of alternative pathways for acetate production in C. tyrobutyricum (Zhu et al. 2004; Liu and Yang 2006). Recently, Charrier et al. (Charrier et al. 2006) discovered a novel CoA transferase pathway in human colonic bacterium Roseburia sp. A2-183, which used acetate: butyryl-CoA CoA- transferase for butyrate formation in the human colon. The CoA transferase has also been located in the 454 sequencing database of C. tyrobutyricum (Chapter 4). It is very likely that this novel CoA-transferase provides alternative pathways for acetate and butyrate biosynthesis in clostridia. Future study will be conducted to investigate the CoA transferase function in C. tyrobutyricum.

Compared to the wild type, the ptb mutant had a higher acetate production, which stimulated the higher ATP formation and resulted in the enhanced specific growth rate in both glucose and xylose fermentations.

The ptb mutant showed an increased tolerance to butyric acid inhibition, which can be partially attributed to the increased production of acetate and ATP. The exact acid tolerance mechanism in C. tyrobutyricum is still unknown, but should be related to the ability of H+-ATPase to pump out the proton from its intracellular space at the expense of

ATP consumption (O'Sullivan and Condon 1999; Cotter and Hill 2003). The increased

75

ATP production in the mutant thus can be used to pump the proton out of the cell and to maintain its suitable pH environment. End product inhibition is one of the limiting factors on the organic acid fermentation. It causes low product titer and hence high production cost. The ptb mutant with increased butyrate tolerance, growth rate, and productivity is thus a promising candidate for industrial fermentation to produce butyric acid and hydrogen from glucose and xylose.

3.5 Conclusion

The ptb-disrupted mutant of C. tyrobutyricum has been constructed and confirmed by PCR and reduced phosphotransbutyrylase and butyrate kinase activities. Compared to the parental strain, the mutant had a higher specific growth rate and significantly lower b/a ratio from both glucose and xylose in batch fermentations. However, butyrate production did not decrease significantly, suggesting the existence of another pathway leading to butyrate production.

76

3.6 References

Andersch, W.;H. BahlG. Gottschalk (1983). "Level of enzymes involved in acetate, butyrate, acetone and butanol formation in Clostridium acetobutylicum." European Journal of Applied Microbiology and Biotechnology 18(6): 327-332.

Boyles, D. (1984). Bioenergy technology thermodynamics and costs. New York, Wiley.

Bruggemann, H.;S. Baumer;W. F. Fricke;A. Wiezer;H. O. Liesegang;I. Decker;C. Herzberg;R. Martinez-Arias;R. Merkl;A. HenneG. Gottschalk (2003). "The genome sequence of Clostridium tetani, the causative agent of tetanus disease." PNAS 100(3): 1316-1321.

Cary, J. W.;D. J. Perterson;E. T. PapoutsakisG. Bennett, N. (1988). "Cloning and expression of Clostridium acetobutylicum phosphotransbutyrylase and butyrate kinase genes in Escherichia coli." J. Bacterial 170: 4613-4618.

Charrier, C.;G. J. Duncan;M. D. Reid;G. J. Rucklidge;D. Henderson;P. Young;V. J. Russell;H. F. AminovP. Louis (2006). "A novel class of CoA-transferase involved in short-chain fatty acid metabolism in butyrate-producing human colonic bacteria." Microbiology 152: 179-185.

Copeland, A.;S. Lucas;A. Lapidus;K. Barry;J. C. Detter;T. Glavina;N. Hammon;S. Israni;E. Dalin;H. Tice;S. Pitluck;D. Sims;T. Brettin;D. Bruce;R. Tapia;J. Brainard;J. Schmutz;F. Larimer;M. Land;L. Hauser;N. Kyrpides;N. Mikhailova;G. Bennet;I. Cann;J. S. Chen;A. L. Contreras;D. Jones;E. Kashket;W. Mitchell;S. Stoddard;W. Schwarz;N. Qureshi;M. Young;Z. Shi;T. Ezeji;B. White;H. BlaschekP. Richardson (2009). "Clostridium beijerinckii NCIMB 8052, complete genome." http://www.ncbi.nlm.nih.gov/nuccore/CP000721.

Cotter, P. D.C. Hill (2003). "Surviving the acid test: response of gram-positive bacteria to low pH." Microb. Molec. Bio. Rev. 67(3): 429-453.

Dziedzak, J. D. (1986). "Biotechnology and flavor development: An industrial research perspective " Food Technology 40: 8-20.

Ellman, G. L. (1959). Arch. Biochem. Biophysics 82(1): 70-77.

Green, E. M.G. N. Bennet (1997). "Genetic manipulation of acid and solvent formation in Clostridium acetobutylicum ATCC824." Biotech. Bioeng. 58: 215-221.

Green, E. M.;Z. L. Boynton;L. M. Harris;F. B. Rudolph;E. T. PapoutsakisG. N. Bennett (1996). "Genetic manipulation of acid formation pathways by gene inactivation in Clostridium acetobutylicum ATCC824." Microbiol 142: 2079-2086.

77

Guerout-Fleury, A. M.;K. Shazand;N. FrandsenP. Stragier (1995). "Antibiotic-resistance cassettes for Bacillus subtilis." Gene 167: 335-336.

Hartimanis, M. G.S. Gatenbeck (1984). "Intermediary metabolism in clostridium acetobutylicum: levels of enzymes envolved in the fermentation of acetate and butyrate." Appl. Environ 47: 1277-1283.

Huang, Y. L.;K. Mann;J. M. NovakS. T. Yang (1998). "Acetic acid production from fructose by Clostridium formicoaceticum immobilized in a fibrous-bed bioreactor." Biotech. Prog. 14: 800-806.

Jiang, L.;J. F. Wang;S. Z. Liang;X. N. Wang;P. L. CenZ. N. Xu (2009). "Butyric acid fermentation in a fibrous bed bioreactor with immoblilized Clostridium tyrobutyricum from cane molasses." Bioresour Technol 100(13): 3403-3409.

Jiraskova, A.;L. Vitek;J. Fevery;T. RumlP. Branny (2005). "Rapid protocol for electroporation of Clostridium perfeingens." J. Microbiol Methods 62: 125-127.

Jo, J. H.;D. S. LeeJ. M. Park (2008). "The effects of pH on carbon material and energy balances in hydrogen-producing Clostridium tyrobutyricum JM1." Bioresour Technol 99(17): 8485-8491.

Jones, D. T.D. R. Woods (1986). "Acetone-butanol fermentation revisited." Microb. Rev. 50(4): 484-524.

Klijn, N.;F. Nieuwenhof;J. Hoolwerf;V. d. W. C.A. Weerkamp (1995). "Identification of Clostridium tyrobutyricum as the causative agent of late blowing in cheese by species- specific PCR amplification." Appl. and Enviro. Microbiol. 61(8): 2919-2924.

Lipmann, F.L. C. Tuttle (1945). "A specific micromethod for the determination of acyl phosphate." J. Bio. Chem. 159: 21-28.

Liu, X.;Y. ZhuS. T. Yang (2006). "Construction and characterization of ack deleted mutant of Clostridium tyrobutyricum for enhanced butyric acid and hydrogen production." Biotech. Prog. 22: 1265-1275.

Liu, X. G.S. T. Yang (2006). "Kinetics of butyric acid fermentation of glucose and xylose by Clostridium tyrobutyricum wild type and mutant." Proc. Biochem. 41(4): 801-808.

Michel-Savin, D.;R. MarchalJ. P. Vandecasteele (1990a). "Butyrate production in continuous culture of Clostridium tyrobutyricum: effect of end-product inhibition." Appl. Microbio. Biotech. 33(2): 127-131.

78

Michel-Savin, D.;R. MarchalJ. P. Vandecasteele (1990b). "Control of the selectivity of butyric acid production and improvement of fermentation performance with Clostridium tyrobutyricum." Appl. Microbio. Biotech. 32(4): 387-392.

Michel-Savin, D.;R. MarchalP. Vandecasteele (1990c). "Butyric fermentation; Metabolic behavior and production performance of Clostridium tyrobutyricum in a continuous culture with cell recycle." Appl. Microbiol. Biotechnol. 34(172-177).

Nolling, J.;G. BretonM. V. Omelchenko (2001). "Genome sequence and comparative analysis of the solvent-producing bacterium Clostridium acetobutylicum." J. Bacteriol 183: 4823-4838.

O'Sullivan, E.S. Condon (1999). "Relationship between acid tolerance, cytoplasmic pH, and ATP and H+-ATPase levels in chemist cultures of Lactococcus lactis." Appl. Environ. Microbiol. 65: 2287-2293.

Pouillart, P. R. (1998). "Role of butyric acid and its derivatives in the treatment of colorectal cancer and homoglobinopathies." Life Sci. 63(20): 1739-1760.

Rose, I. A. (1955). "Acetate kinase of bacteria (acetokinase)." Methods Enzymol. 1: 591- 595.

Sebaihia, M.;M. W. Peck;N. P. Minton;N. R. Thomson;M. T. Holden;W. J. Mitchell;A. T. Carter;S. D. Bentley;D. R. Mason;L. Crossman;C. J. Paul;A. Ivens;M. H. Wells- Bennik;I. J. Davis;A. M. Cerdeno-Tarraga;C. Churcher;M. A. Quail;T. F. Chillingworth;T. Feltwell;A. Fraser;I. Goodhead;Z. Hance;K. Jagels;N. Larke;M. Maddison;S. Moule;K. Mungall;H. Norbertczak;E. Rabbinowitsch;M. Sanders;M. Simmonds;B. WhiteS. P. Whithead, J. (2007). "Genome sequence of a proteolytic (Group I) Clostridium botulinum strain Hall A and comparative analysis of the clostridial genomes." Genome Res. 17(1082-1092).

Sebaihia, M.;B. W. Wren;P. Mullany;N. F. Fairweather;N. Minton;R. Stabler;N. R. Thomson;A. P. Roberts;A. M. Cerdeno-Tarraga;H. W. Wang;M. T. Holden;A. Wright;C. Churcher;M. A. Quail;S. Baker;N. Bason;K. Brooks;T. Chillingworth;A. Cronin;P. Davis;L. Dowd;A. Fraser;T. Feltwell;Z. Hance;S. Holroyd;K. Jagels;S. Moule;K. Mungall;C. Price;E. Rabbinowitsch;S. Sharp;M. Simmonds;K. Stevens;L. Unwin;S. Whithead;B. Dupuy;G. Dougan;B. BarrellJ. Parkhill (2006). "The multidrug-resistant human pathogen Clostridium difficile has a highly mobile, mosaic genome." Nature Gene. 38: 779-786.

Shimizu, T.;K. Ohtani;H. Hirakawa;K. Ohshima;A. Yamashita;T. Shiba;N. Ogasawara;M. Hattori;S. KuharaH. Hayashi (2002). "Complete genome sequence of Clostridium perfringens, an anaerobic flesh-eater." PNAS 99(2): 996-1001.

79

Valentine, R. C.R. S. Wolfe (1960). "Purification and role of phosphotransbutyrylase." J. Bio. Chem. 235: 1948-1952.

Walter, K.;R. Nair;J. CaryG. Bennett (1993). "Sequence and arrangement of two genes of the butyrate-synthesis pathway of Clostridium acetobutylicum ATCC824." Gene 134: 107-111.

Williams, E. A.;J. M. CoxheadJ. C. Mathers (2003). "Anti-cancer effects of butyrate: use of microarray technology to investigate mechanisms." Proc. Nutr. Soc. 62: 107-115.

Wisenborn, D. P.;F. B. RudolphE. T. Papoutsakis (1989). "Phosphotransbutyrylase from Clostridium acetobutylicum ATCC824 and its role in acidogenesis." Appl. and Enviro. Microbiol. 66: 317-322.

Wu, Z. T.S. T. Yang (2003). "Extractive fermentation for butyric acid production from glucose by Clostridium tyrobutyricum." Biotech. Bioeng. 82: 93-102.

Zhu, Y.;X. LiuS. T. Yang (2004). "Construction and characterization of pta gene-deleted mutant of Clostridium tyrobutyricum for enhanced butyric acid fermentation." Biotech. Bioeng. 90: 154-166.

Zhu, Y.;Z. T. WuY. S.T. (2002). "Butyric acid production from acid hydrolysate of corn fiber by Clostridium tyrobutyricum in a fibrous-bed bioreactor." Proc. Biochem. 38(5): 657-666.

Zhu, Y.S. T. Yang (2003). "Adaptation of Clostridium tyrobutylicum for enhanced tolerance to butyric acid in a fibrous-bed bioreactor." Biotech. Prog. 19: 365-372.

Zigova, J.;E. Sturdisk;D. VandakS. Schlosser (1999). "Butyric acid production by Clostridium butyricum with integrated extraction and pertraction." Proc. Biochem. 34: 153-160.

80

Strains / Plasmids Characteristics Source or reference

Clostridium tyrobutyricum ATCC

ATCC 25755

E. coli DH5α Invitrogen pCR4-TOPO Apr, Kanr Invitrogen pDG647 Emr, Apr Guerout-Fleury et al. 1995 pCR4ptb Apr, Kanr, ptb fragment This study pPtbEm Emr, Apr, Kanr, ptb This study

fragment

Table 3.1 Strains and plasmids used in this study.

81

Glucose + 2NAD 2ADP

2NADH 2ATP 2 Pyruvate

Fd 2H+ Fd NAD+ pfor hyd hyd nfor CO H NADH 2 FdH2 2 FdH2 2 Acetate ack pta 2 Acetyl‐CoA

ATP ADP Thiolase Acetoacetyl‐CoA NADH hbd NAD+ 3‐hydroxybutyryl CoA crotonase Crotonyl CoA NADH Butyryl CoA dehydrogenase buk ptb NAD+ Butyrate Butyryl‐CoA

ATP ADP

Figure 3.1 The metabolic pathway showing butyric acid, acetic acid, and hydrogen production from glucose in C. tyrobutyricum.

82

Figure 3.2 Alignment of amino acid sequences of PTB of several Clostridium species showing highly conserved regions. Arrows indicate the conserved sequences used in designing the forward and reverse primers for PCR amplification of ptb from C. tyrobutyricum.

83

C: Clostridium specific primers M: DNA marker

1: C. acetobutylicum specific primers 2: C. tyrobutyricum specific primers M: DNA marker

Figure 3.3 The nested-PCR to confirm the chromosome from C. tyrobutyricum (3a: Clostridium specific primers; 3b: species specific primers)

84

Figure 3.4 Construction of the non-replicative integrational plasmid pPtbEm with one 0.64-kb ptb fragment cloned from C. tyrobutyricum. Abbreviations: ptb, partial ptb gene; Apr, ampicillin resistance gene; Emr, erythromycin resistance gene; Kanr: kanamycin resistance gene; pUC ori: G(-) replication origin.

85

ptb WT Chromosome DNA

Index

1.6 kb Emr 2.2 kb Emr inserted ptb fragment Non-replicative plasmid pPtbEm ptb gene in chromosome

Homologous recombination

Inactivated ptb in chromosome Inactivated ptb in chromosome

Figure 3.5 Homologous recombination of pPtbEm containing an Emr-inserted ptb fragment with the chromosome of wild type resulting in the disruption of the ptb gene.

86

Figure 3.6 PCR characterization of the ptb mutant. A) using ptb primers; B) using Emr primers.

87

140% wt ptb 120%

100%

80%

60%

40% Relative Activity Relative

20%

0% PTA PTB AK BK

Figure 3.7 Comparison of activities of PTA, PTB, AK, and BK in wild type (wt) and ptb mutant. (PTA: phosphotransacetylase; PTB: phosphotransbutyrylase; AK: acetate kinase; BK: butyrate kinase)

88

3.5

3 ratio

B/A 2.5

or 2 (g/l) butyrate 1.5 acetate 1 B/A (g) 0.5 Concentration 0 WT PTB‐1PTB‐2PTB‐3PTB‐4PTB‐5

Figure 3.8 Effect of ptb-disruption on B/A ratio in batch fermentation of glucose by C. tyrobutyricum wild type and ptb mutants.

89

5 WT 4

3 600

OD 2

1

0 0 10203040 Time (h) 5

4 ptb 3 600

OD 2

1

0 0 10203040 Time (h) 0 g/l 5 g/l 10 g/l 15 g/l

120%

100%

80% Kp = 13.5

60%

40% Kp = 9.0

Relative Growth Rate Relative 20% wt ptb 0% 0481216 Butyrate (g/L)

Figure 3.9 Noncompetitive inhibition of butyric acid on cell growth of C. tyrobutyricum wild type (wt) and ptb mutant (ptb).

90

CHAPTER 4

KINETICS OF BUTYRIC ACID FERMENTATION BY CLOSTRIDIUM

TYROBUTYRICUM PTB MUTANT

Summary

Fermentation kinetics of glucose and xylose by Clostridium tyrobutyricum wild

type and ptb mutant were studied. A simplified growth medium without vitamins and

trace metals supplementation was developed based on a partial factorial experimental

design and statistical analysis of the fermentation results. The optimal pH was found to

be 6 and the optimal substrate concentration was 50 g/l by using one-factor-a-time

screening. The impact of ptb-disruption on the fermentation kinetics was studied using

glucose and xylose as substrates. Compared to the wild type, the B/A ratio decreased by

34.7% for glucose and 28.1% for xylose, which could be attributed to the inactivation of

ptb. The productivity of hydrogen increased 38% for glucose and 46% for xylose, which

could be partially attributed to the higher specific growth rate (33% increase for glucose

and 40% increase for xylose). The butyric acid fermentation and hydrogen production

were further improved by immobilizing cells in the fibrous-bed bioreactor to facilitate

cell adaptation to attain a higher final organic acid concentration.

91

4.1 Introduction

Plant biomass is an abundant renewable energy resource and constantly being

created by photosynthesis. A typical plant biomass has about 25% lignin and 75% cellulose containing glucose and hemicellulose containing glucose and xylose. The

cellulose and hemicellulose can be broken down by acid or enzymatic hydrolysis to

release its monomer glucose as fermentable sugar (Aristidou and Penttila 2000). While

hydrolysis of cellulose has been greatly improved recently with new technologies in

pretreatment and enzyme modification, many of current industrial microbial

fermentations cannot use pentoses (e.g., xylose) as the substrate. C. tyrobutyricum has been demonstrated for its ability to convert xylose to butyric acid at the same rate comparable to glucose (Zhu et al. 2002).

Besides lignincellulose, bio-wastes, such as whey permeate and glycerol, are also regarded as economical carbon sources for bio-based chemicals and biofuels production.

Whey permeate is the byproduct of whey protein production and has been used for lactic acid fermentation (Tejavadi and Cheryan 1988; Norton et al. 1994). Glycerol is the

byproduct from biodiesel production and has been used for production of ethanol,

succinic acid and 1,3-propanediol especially because of its highly reducing status (Ito et

al. 2005; Dharmadi et al. 2006; Wang et al. 2007).

Medium composition affects the economics of a fermentation process. The major

components of microbial fermentation medium include nitrogen sources, carbon sources

92

and buffers to control the pH value, ionic strength, and redox potential. The composition

of the basal medium for C. tyrobutyricum is complicated and contains 27 components. As

it contains expensive vitamins, the medium is uneconomical and its preparation is labor

intensive. Traditional optimization method is one-factor-a-time by which one factor is

optimized while the others are fixed. This single-dimensional optimization was time-

consuming and not able to reach the optimal conditions because it ignores the interaction

among variables (Chen and Lee 1997), especially when many factors need to be

considered. Fractional factorial design was employed for screening multifactors to find

the most significant independent factors at lower run numbers. In this work, seven medium components in the Clostridium Growth Medium (CGM) were screened to

simplify the composition of the medium and free-cell fermentation was performed to

verify that there was no negative effect using the simplified medium.

The initial sugar concentration and pH affect cell growth and product yield,

productivity and titer. pH affects the metabolic pathway by changing the activities of key

enzymes (Tanisho et al. 1987; Dabrock et al. 1992.; Lay 2000). It has been reported that pH could impact the ionization state of enzymes (Fabiano and Perego 2002). In the case

of organic acid production, a high initial pH value can slow down the pH drop and thus

alleviate proton inhibition. A lower pH is helpful to separate organic acid from the

fermentation broth by extractive fermentation (Zigova et al. 1999; Wu 2003; Wu and

Yang 2003). It has also been noticed that a low initial pH also led to a long lag phase

(Khanal et al. 2004). On the other hand, high initial pH values reduced the length of the

93

lag phase, but led to a lower yield of hydrogen production (Zhang et al. 2003). The pH

range for Clostridia growth has been determined to be 4.5-7.0 (Zigova et al. 1999).

The main problem associated with conventional carboxylic acid fermentation is

the product inhibition, which causes low product concentration, and consequently, low

yield and low productivity, resulting in high costs in product recovery (Michel-Savin et al.

1990). This problem can be partially addressed by using the fibrous-bed bioreactor (FBB)

where bacteria are immobilized in a fibrous matrix to achieve a high cell density (US

patent no. 5,563,069). This method has been successfully employed in the production of

organic acids, including lactic acid, acetic acid, propionic acid, and butyric acid (Yang et

al. 1994; Silva and Yang 1995; Yang et al. 1995; Yang 1996; Huang et al. 1998; Huang

and Yang 1998; Zhu and Yang 2003).

In this work, CGM was simplified using fractional factorial experimental design.

Different substrates were tested for the fermentation by the ptb mutant of C.

tyrobutyricum and the fermentation conditions were optimized. The kinetics of free-cell

batch fermentations was investigated using glucose and xylose as substrates to study the

impact of ptb-disruption on the fermentation profile. Furthermore, cells were adapted by

immobilization in the fibrous-bed bioreactor and higher final product concentrations were

obtained in the FBB fermentation compared to free-cell fermentations.

4.2 Materials and methods

4.2.1 Culture and media

94

The stock culture of C. tyrobutyricum ATCC 25755 was stored in

Difco™ Reinforced Clostridia Medium (RCM) in serum bottles under anaerobic

conditions at 4 °C. Unless otherwise noted, the medium used for fermentation studies was

the Clostridia Growth Medium (CGM) as previously described by Huang et al (Huang et

al. 1998) and supplemented with glucose or xylose as the carbon source.

A Simplified Clostridia Growth Medium (SCGM) described later, was also developed

and used in this study. The basal medium and the concentrated sugar solution, i.e. 500 g/l

glucose or xylose, were autoclaved separately at 121°C and 15 psi for 30 min to avoid

undesirable reactions. The concentrated sugar solution was then pumped into the

fermenter aseptically.

4.2.2 Design of experiments for simplifying medium composition

JMP 7.0 was used to design the factor screening experiment. Seven factors were

screened. Among them, Fe2+ solution, Ni2+ solution, vitamins solution and trace metals

solution were categorical factors, while yeast extract, TrypticaseTM soy broth, and

Trypton were 2-level numerical factors, whose values were limited between 0 and 5 g/l. 2 center points and 1 duplicate were added and the total run number was 36 with a

resolution of 4 as shown in Table 4.1. The experiments were performed in serum tubes

each containing 10 ml of media with 10g/l glucose. Three responses which were cell

growth, butyric acid final concentration and total gas production were measured. The

OD600 of each serum tube culture was monitored directly without sampling by using a

95

modified cuvette holder and a spectrophotometer (Sequoia-Turner, model 340). The total

gas was collected and the gas volume was measured. The final butyric acid concentration was measured by HPLC

4.2.3 Effect of initial glucose concentration

The impact of initial glucose concentration on cell growth was evaluated in serum tubes each containing 10 ml RCM supplemented with various amounts of glucose to make the final glucose concentration in the range of 10 g/l – 90 g/l. Water was added to make the final volume to 12 ml. The culture was inoculated with 0.5 ml fresh inoculum.

The OD600 was measured with a spectrophotometer at 600 nm at regular time intervals.

The specific growth rate was calculated based on the slope of the semi-logarithmic plot of

OD600 vs. time during the exponential phase.

4.2.4 Effect of pH

The pH influence on cell growth was evaluated in serum tubes each containing 10

ml RCM. The medium pH was adjusted to 4 ~ 9 by adding 1 N HCl or 1 N NaOH. The

small difference in the liquid volume was ignored. A 0.5 ml fresh culture in the

exponential phase was added into each tube as inoculum. The OD600 was measured with a

modified spectrophotometer at 600 nm at regular time intervals. The specific growth rate

96

was calculated based on the slope of the semi-logarithmic plot of OD600 vs. time during

the exponential phase.

4.2.5 Free-cell fermentations

The kinetics of free-cell fermentations of wild type and ptb-disrupted C.

tyrobutyricum were studied in 2-liter stirred bioreactor containing 1 l CGM supplemented

with 50 g/l glucose or xylose. The sugar substrate was prepared in concentrated solutions

(500 g/l) and pumped into the fermenter aseptically. The medium was sparged with

filtered nitrogen gas for 2 hours to achieve anaerobiosis. Each reactor was inoculated

with 50 ml of a fresh cell suspension in the exponential phase (OD600 = 2.0) grown in

serum bottles. Batch fermentations were performed at 37ºC and pH 6.0 ± 0.1 adjusted

with 6 N NaOH. CO2 and hydrogen production was monitored by the water replacement

method along with GC analysis. Liquid samples were taken at regular intervals (8-12 h)

to analyze the cell density and concentrations of the substrate and products in the

fermentation broth. For fed-batch fermentations, a concentrated substrate solution was

added periodically when the glucose concentration was close to 0 g/l.

4.2.6 Immobilized-cell fermentations in FBB

The FBB fermentation system (Figure 4.1) consisted of a 5-l stirred-tank

fermenter (Marubishi MD-300) and a 500 ml fibrous-bed bioreactor (FBB). The FBB

97

was made of a glass column with a water jacket to control the temperature. The inside of

the bioreactor was packed with a spirally wound cotton towel supported by a stainless

steel wire mesh. The working volume of the bioreactor was 500 ml. The FBB and the stirred-tank fermenter were connected to each other through a recirculation loop. The

FBB system was operated under anaerobic conditions at pH 6 and 37°C. The anaerobic

condition was maintained through N2 sparging until gas was produced.

To start the fermentation, 100 ml of a fresh cell suspension were inoculated into the fermenter containing 2 l CGM and then grown at 37°C, 200 rpm and pH 6, which was controlled by adding 6N NaOH. The culture medium was circulated at 10 ml/min when cell optical density (OD600) reached 5.0. The slow circulation rate allowed cells to attach

to the matrix and to be stabilized. After 2 days of circulation, cells were mostly

immobilized in the matrix and the OD600 was stable again (~10). At this time, the

circulation rate was increased to 50 ml/min for a better mixing. Additional carbon source

was supplied when the glucose concentration was close to zero. The glucose feeding was

repeated several times to allow the fermentation to continue and to increase the butyrate

concentration until no glucose was consumed.

4.2.7 Analytical methods

The cell density was calculated based on OD600 measured with a

spectrophotometer (Sequoia-Turner, model 340). One unit of OD600 was corresponding to

0.68 g/l of cell dry weight.

98

The composition of gas produced in the fermentation was analyzed with a gas

chromatograph (Shimadzu GC 2014) equipped with Thermo Conductive Detector and

RT-QPLOT column (Restek) with argon as the carrier gas at the flow rate of 4.38 ml/min.

The injector temperature was 60°C, the column temperature was 30°C and the detector

temperature was 60°C. The volume of produced gas was measured by the water

replacement method.

High performance liquid chromatography (HPLC) with an organic acid analysis

column (Aminex HPX-87H, Biorad) was used to analyze sugar and organic acid contents,

including xylose, glucose, acetate, and butyrate in fermentation broth samples. The

detailed method has been described in appendix C (Zhu and Yang 2003).

4.3 Results

4.3.1 Simplified medium composition

The medium composition plays a critical role in the economics of bio-based

chemicals production. A medium using inexpensive materials and simplified preparation

process can decrease the production cost. The basal Clostridium Growth Medium (CGM)

used for C. tyrobutyricum fermentation consisted of the following components (per liter):

a) 40 ml of mineral #1 solution containing 7.86 g/l K2HPO4·3H2O.

b) 40 ml of mineral #2 solution consisting of (per liter): 6 g KH2PO4; 6 g (NH4)2SO4; 12

g NaCl; 2.5 g MgSO4·7H2O; 0.16 g CaCl2·2H2O.

99

c) 10 ml of the trace metals solution containing (per liter): 1.5 g nitrilotriacetic acid; 0.1

g FeSO4·7H2O; 0.5 g MnSO4·2H2O; 1.0 g NaCl; 0.1 g CoCl2; 0.1 g CaCl2·2H2O; 0.1

g ZnSO4·5H2O; 0.01 g CuSO4·5H2O; 0.01 g AlK(SO4)2; 0.01 g H3BO3; 0.01 g

Na2MoO4·3H2O.

d) 10 ml of the vitamin solution containing (per liter): 5 mg thiamine-HCl; 5 mg

riboflavin; 5 mg nicotinic acid; 5 mg pantothenate; 0.1 mg vitamin B12; 5 mg p-

aminobenzoic acid; 5 mg lipoic acid.

e) 10 ml of 0.005% (w/v) NiCl2·6H2O solution.

f) 2 ml of 0.2% (w/v) FeSO4·7H2O solution.

g) 5 g TrypticaseTM soy broth (BD, catalog # 211768). h) 5 g yeast extract.

The preparation of the CGM was time-consuming and labor intensive. As yeast

extract and TrypticaseTM soy broth, containing casein and soybean protein hydrolysates,

were used as the nitrogen source, they could also provide vitamins and trace metals that

are critical to cell growth, butyric acid production and hydrogen production. Furthermore, another nitrogen source Trypton containing casein hydrolysate was also tested.

A ¼ fractional factorial design with 1 duplicate and 2 center points was used to examine if the vitamin solution, trace metals solution, NiCl2 solution, FeSO4 solution, yeast extract, TrypticaseTM soy broth, and Trypton are significant factors affecting C.

tyrobutyricum ptb mutant fermentation. The result (figure 3.2) showed that only yeast

extract and soy broth were significant factors with p = 0.0015 and p = 0.0016 for butyric

100

acid production and p = 0.0010 and p = 0.0001 for cell growth, respectively. In addition

to these two nitrogen sources, the NiCl2 solution showed a significant effect on hydrogen

production with p = 0.0069, suggesting that the [NiFe]-hydrogenase was active in C.

tyrobutyricum. It is well known that Fe plays a critical role in hydrogenase activity.

FeSO4 was thus included in the simplified medium although it did not show significant

impact on any of the three responses (p=0.8915, 0.7458, 0.5971 for cell growth, hydrogen

production and butyric acid production, respectively) due to factors interaction probably.

It was apparent that the trace metals solution and the vitamin solution had no significant

effects on the fermentation and thus could be excluded from the CGM.

Based on these findings, the medium was simplified to consist of only yeast

extract (5 g/l), Trypton (5 g/l) and the following two solutions (per liter):

a) 40 ml of mineral #1 solution containing 7.86 g/l K2HPO4·3H2O

b) 40 ml of mineral #2 solution consisting of (per liter): 6 g KH2PO4; 6 g (NH4)2SO4; 12 g NaCl; 2.5 g MgSO4·7H2O; 0.16 g CaCl2·2H2O; 0.125 g NiCl2 and 0.1 g FeSO4 (In order

to prevent Fe2+ oxidation and precipitation during autoclaving, the solution mineral #2

was sterilized by microfiltration).

Free-cell batch fermentation was performed to verify that the simplified medium

did not have any negative impact on cell growth, butyric acid and hydrogen production.

As shown in Figure 4.3, both the original and the modified medium gave similar specific growth rates (0.20 h-1 vs. 0.21 h-1) indicating that medium simplification did not change

101

cell growth. Similar results were also found for butyric acid yield (0.74 vs. 0.73) and hydrogen yield (1.6 vs. 1.62).

4.3.2 Substrate selection

The fermentation substrate is critical to the economy of bio-based chemicals production. Generally, low-cost substrates can be derived from agricultural residues containing starch, cellulose and hemicellulose, or from industrial wastes, such as lactose in whey permeate and glycerol in biodiesel wastewater. As shown in Figure 4.4, C. tyrobutyricum does not have the ability to use lactose and glycerol, although these two substrates have been used for propionic acid production and other fermentation processes.

However, C. tyrobutyricum can uptake xylose, a major component of hemicellulose consisting of around 20% of the dry weight of renewable biomass. However, the specific

growth rate and the product yield from xylose were slightly lower than those from

glucose.

4.3.3 Optimized fermentation conditions

A high glucose concentration caused longer lag phase (Figure 4.5a) and inhibit

cell growth (Figure 4.5b). In general, no significant effect on cell growth was observed

when the glucose concentration was less than 70 g/l. At 70 g/l, the fermentation showed a

longer lag phase, although the specific growth rate was not affected obviously. Further

102

increasing glucose concentration to 90 g/l showed significant inhibition on cell growth with p = 0.027. Considering the long lag phase would decrease productivity dramatically,

50 g/l of glucose appeared to be the optimal initial glucose concentration for the batch fermentation. Since xylose showed a similar metabolic rate as glucose (Zhu and Yang

2003), 50 g/l was also considered as the optimal initial concentration for xylose.

The pH of the medium is another important factor in fermentation affecting the enzyme activity by changing the ionization state of enzyme’s components (Fabiano and

Perego 2002). Consequently, the pH affects the carbon source distribution (Zhu et al.

2004). For C. tyrobutyricum, the highest specific growth rate was observed at pH 6, which also had the shortest lag phase (Figure 4.6).

4.3.4 Effects of ptb-disruption on fermentation kinetics

Free cell batch fermentations with the wild type and ptb mutant were performed at pH 6.0 and 37 °C to examine the effects of ptb-disruption on the fermentation kinetics with glucose and xylose as the substrate. The batch fermentation results are shown in

Figures 4.7 and 4.8, and the kinetic data are summarized in Table 4.2. Compared to the wild type, the B/A ratio of ptb mutant fermentation decreased 34.7% for glucose and 28.1% for xylose, which can be attributed to the inactivation of ptb. However the butyrate yield was still comparable to that of the wild type, suggesting a new pathway other than PTB-

BK route could also produce butyrate. Charrier et al. (Charrier et al. 2006) discovered a novel CoA transferase pathway in human colonic bacterium Roseburia sp. A2-183,

103

which used butyryl-CoA: acetate CoA-transferase for butyrate formation in the human

colon. It is very likely that this butyrate biosynthesis pathway is also used in C.

tyrobutyricum since CoA transferase genes were also found in the 454 genome sequences

of this bacterium.

The hydrogen yield by the ptb mutant increased 14% for glucose and 17% for

xylose, supporting that the lower B/A ratio accompanied a higher hydrogen yield. In

addition, the productivity of hydrogen increased 38% for glucose and 46% for xylose

(Table 4.2), which could be partially attributed to a higher specific growth rate (33%

increase for glucose and 40% increase for xylose). As shown in the metabolic pathway of

C. tyrobutyricum, 2 mol of ATP are produced from 1 mol glucose if acetate is the only

end product, while only 1 mol ATP is produced with butyrate formation. The increased

acetate production in the ptb mutant caused more ATP production, which benefited cell

growth and thus hydrogen productivity.

4.3.5 Free cell fed-batch fermentation kinetics

Figure 4.9 shows the fed-batch fermentation kinetics with free cells, which was

performed to identify the butyrate inhibition level for the ptb mutant. In the first batch,

50 g/l glucose was completely consumed in 65 h and 16.6 g/l butyric acid and 9900 ml of

hydrogen were produced. The butyric acid yield was 0.70 (mol/mol), the acetic acid yield

was 0.137 (mol/mol), and the hydrogen yield was 1.67 (mol/mol). The second batch of 50

g/l glucose was added when the glucose concentration was close to zero and it was

consumed at a slower rate due to the butyrate inhibition. The bacteria stopped producing

104

butyric acid and hydrogen when the butyric acid concentration reached 21.3 g/l, which

was higher than that for the wild type (16.3 g/l) reported by Zhu and Yang (Zhu et al.

2002). This finding is consistent with the results from serum bottle fermentations

(Chapter 3.3.4), which showed that the ptb mutant had a higher butyrate tolerance than the wild type with Kp = 13.5 g/l vs. Kp = 9.0 g/l. In the second batch, the butyrate yield was lower than the first batch due to the final product inhibition, which led to an overall butyric acid yield of 0.59 (mol/mol). The final (overall) hydrogen yield from the fed- batch fermentation was 1.68 mol/mol glucose.

4.3.6 FBB fed-batch fermentation kinetics

The maximum butyric acid concentration could be produced by the ptb mutant was also evaluated with cells immobilized and adapted to tolerate a higher butyrate

concentration in fed-batch fermentation. Figure 4.10 shows the fermentation kinetics with glucose at pH 6.0 in the FBB by the ptb mutant. As shown in table 4.3, the maximum butyrate concentration was increased to 31.5 g/l, as compared to 21.3 g/l in the free-cell fermentation. Also noticed is the decreased B/A ratio (1.95 vs. 2.7), which again accompanied a higher hydrogen production (1.90 vs. 1.68 mol/mol).

4.4 Discussion

Generally, hydrogen production occurs in the transfer of reducing equivalents

105

produced during the pyruvate oxidation using the pyruvate-ferredoxin oxidoreductase

(PFOR) coupling with the hydrogenase (Gray and Gest 1975; Cammack et al. 2001;

Hallenbeck and Benemann 2002), as illustrated by the following equations:

Pyruvate + CoA + Ferredoxin (ox) → Acetyl-CoA + CO2 + Ferredoxin (red)

Ferredoxin (red) → Ferredoxin (ox) + H2

The theoretical molar yield of hydrogen is two per mol glucose if the reducing

equivalent from pyruvate is the only reducing source of the hydrogen. However, some

clostridia strains also contain NADH: ferredoxin oxidoreductase (NFPR), which can

produce additional hydrogen using the reducing equivalent stored in NADH during the

glycolysis (Saint-Amans et al. 2001):

NADH + Ferredoxin (ox) → NAD+ + Ferredoxin (red)

Ferredoxin (red) → Ferredoxin (ox) + H2

Although the reaction from NADH to hydrogen is energetically unfavorable under

standard conditions (Hallenbeck 2005), it could be compensated by coupling with the

ATP hydrolysis, which is highly energetically favorable. Compared to the wild type, the

ptb mutant had a higher acetate production accompanied with a higher ATP production to

facilitate the hydrogen production. The higher ATP yield can also explain the increased

specific growth rate of ptb mutant.

It was noticed that in the ptb mutant fermentation, butyrate was still produced at a significant level as compared to the wild type, suggesting the existence of a new butyrate

106

formation pathway other than PTB-BK route. Charrier et al. (Charrier et al. 2006)

discovered a novel CoA transferase pathway in human colonic bacterium Roseburia sp.

A2-183, which used butyryl-CoA: acetate CoA-transferase for butyrate formation in the

human colon. The genome of C. tyrobutyricum was sequenced with 454 technology and the results showed the existence of two genes for acetoacetylCoA: acetate/butyrate CoA

transferase (EC# 2.8.3.9), one of them was located in the contig 64 starting at 129 bp and

ending at 1682 bp and the other one was from 21339 bp to 22931 bp in the contig 22. The deduced amino acid sequences from these two genes show 57% identity to each other.

The one located in contig 64 shows 80% identity to the CoA transferase from Clostridium

Novyi (Gi: 4541513) and the other one shows 76% to the same protein. These CoA transferases may be responsible for transferring the CoA group from butyryl-CoA to acetate to form butyrate and acetyl-CoA. It is thus very likely that C. tyrobutyricum possesses both PTB-BK and CoA transferase pathways for butyrate biosynthesis, explaining why butyrate production was not eliminated in the ptb mutant. Similarly, the existence of CoA transferase also can explain acetate production in pta/ack knock-out mutants (Zhu et al. 2004; Liu et al. 2006).

It has been proposed that weak organic acids inhibit cell growth because they can disturb the pH gradient across the cell membrane (Gu et al. 1998). The mechanism of short-chain organic acid inhibition was that the undissociated acid form of the organic acid under the acidic pH condition in the extracellular space passes the cell membrane due to its non-polar property, and is then dissociated in the neutral environment in the intracellular space, releasing the proton ion and causing pH drop in the intracellular space.

107

In order to control the intracellular pH, extra protons have to be pumped out from the

intracellular space by H+-ATPase at the expense of ATP. In other words, with the same

concentration of butyric acid, a lower pH value will result in more protonated butyric

acid and consumes more ATP to maintain the pH environment in the cells. For C.

tyrobutyricum, it has been noticed that the H+-ATPase performance was related to the butyric acid tolerance. The adapted mutant with higher butyrate tolerance also showed a higher H+-ATPase activity (Zhu and Yang 2003). H+-ATPase is a membrane-bound enzyme essential for maintaining the transmembrane proton motive force required for ion regulation and transportation of certain nutrients (Riebeling and Jungermann 1976; Ivey and Ljungdahl 1986). The higher activity of H+-ATPase can assure that more protons in

the cytoplasm could be transported to the outside of cells, thus maintaining the proper pH

in the cytoplasm. Additionally, the fatty acid composition of the cell membrane also

changed in the adapted cells, which had more saturated long-chain fatty acids in their

membrane, resulting in less membrane fluidity (Zhu and Yang 2003). Similar findings

have also been reported for S. cerevisiae (Casey and Ingledew 1986) and E. coli (Ingram

1976), which have the ability to regulate membrane lipid composition to increase their

tolerance to organic solvents.

4.5 Conclusion

Experimental design with statistical analysis was used to screen significant

components of CGM. The preparation of the CGM was simplified and the medium cost

108

was decreased by removing vitamins and trace metals from the medium. Using one-

factor-a-time screening, the optimal pH was found to be pH 6 and the optimal substrate

concentration was 50 g/l for C. tyrobutyricum batch fermentation. The effects of ptb-

disruption on the fermentation profile were studied using glucose and xylose as

substrates. Compared to the wild type, the B/A ratio decreased by 34.7% for glucose and

28.1% for xylose, which could be attributed to the inactivation of ptb. In addition, the

productivity of hydrogen increased 38% for glucose and 46% for xylose, which could be

partially attributed to the higher specific growth rate (33% increase for glucose and 40%

increase for xylose). Butyric acid and hydrogen production was further enhanced by

immobilizing cells in the FBB to adapt cells to tolerate a higher butyrate tolerance, thus

achieving a higher final product concentration in the fed-batch fermentation.

109

4.6 References

Aristidou, A.M. I. Penttila (2000). "Metabolic engineering applicationto renewable resource utilization." Curr. Opin. Biotech. 11: 187-198.

Cammack, R.;M. FreyR. Robson (2001). Hydrogen as a fuel: Learning from future. London, Taylor & Francis.

Casey, G. P.W. E. Ingledew (1986). "Ethanol tolerance in yeasts." Crit. Rev. Microbiol 13: 219-280.

Charrier, C.;G. J. Duncan;M. D. Reid;G. J. Rucklidge;D. Henderson;P. Young;V. J. Russell;H. F. AminovP. Louis (2006). "A novel class of CoA-transferase involved in short-chain fatty acid metabolism in butyrate-producing human colonic bacteria." Microbiology 152: 179-185.

Chen, W. S.S. L. Lee (1997). "Optimization of medium composition for the production of glucosyltransferase by Aspergillus niger with response surface methodology." Enzyme Microb. Technol., 21(6): 436-440.

Dabrock, B.;H. BahlG. Gottschalk (1992.). "Parameters affecting solvent production by Clostridium pasteurianum. ." Appl. Environ. Microbiol. 164: 35.

Dharmadi, Y.;A. MurarkaR. Gonzalez (2006). "Anaerobic fermentation of glycerol by Escherichia coli: a new platform for metabolic engineering." Biotech. Bioeng. 94(5): 821-829.

Fabiano, B.P. Perego (2002). "Thermeodynamic study and optimization of hydrogen production by Enterobacter aerogenes." Int. J. Hydrog. Energy 27: 149-156.

Gray, C. T.H. Gest (1975). "Biological formation of molecular hydrogen." Science 148: 186-192.

Gu, Z.;A. B. GlatzE. C. Glatz (1998). "Effects of propionic acid on propionibacteria fermentation." Enzyme Microb. Technol. 22: 13-18.

Hallenbeck, P. C. (2005). "Fundamentals of the fermentative production of hydrogen." Water Science & Technology 52: 21-29.

Hallenbeck, P. C.J. R. Benemann (2002). "Biological hydrogen production: fundamentals and limiting processes." Int. J. Hydrogen Energy 27: 1185-1193.

110

Huang, Y. L.;K. Mann;J. M. NovakS. T. Yang (1998). "Acetic acid production from fructose by Clostridium formicoaceticum immobilized in a fibrous-bed bioreactor." Biotech. Prog. 14: 800-806.

Huang, Y. L.S. T. Yang (1998). "Acetate production from whey lactose using co- inmmobilized cells of homolactic and homoacetic bacteria in a fibrous-bed bioreactor." Biotech. Bioeng. 60: 499-507.

Ingram, L. O. (1976). "Adaptation of membrane lipids to alcohols." J. Bacterial 125: 670- 678.

Ito, T.;Y. Nakashimada;K. Senba;T. MatsuiN. Nishio (2005). "Hydrogen and ethanol production from glecerol-containing wastes discharged after biodiesel manufacturing process." J. Biosci. Bioeng. 100(3): 260-265.

Ivey, D. M.L. G. Ljungdahl (1986). "Purification and characteristics of the F1-ATPase from Clostridium thermoaceticum." J. Bacterial 165: 252-257.

Khanal, S. K.;W. H. Chen;L. LiS. W. Sung (2004). "Biological hydrogen production: effects of pH and intermediate products." Int. J. Hydrog. Energy 29 1123-1134.

Lay, J. J. ( 2000). "Modeling and optimization of anaerobic digested sludge converting starch to hydrogen. ." Biotechnol. Bioeng. 68, : 269-278.

Liu, X.;Y. ZhuS. T. Yang (2006). "Construction and characterization of ack deleted mutant of Clostridium tyrobutyricum for enhanced butyric acid and hydrogen production." Biotech. Prog. 22: 1265-1275.

Michel-Savin, D.;R. MarchalP. Vandecasteele (1990). "Butyric fermentation; Metabolic behavior and production performance of Clostridium tyrobutyricum in a continuous culture with cell recycle." Appl. Microbiol. Biotechnol. 34(172-177).

Norton, S.;C. LacroixJ. C. Vuillemard (1994). "Kinetic study of continuous whey permeate fermentation by immobilized Lactobacillus helveticus for lactic acid production." Enzyme Microb. Technol. 16(6): 457-466.

Riebeling, V.K. Jungermann (1976). "Properties and function of Clostridial membrane ATPase." Biochem. Biophys. Acta 430(3): 434-444.

Saint-Amans, S.;L. Girbal;J. Andrade;K. AhrensP. Soucaolle (2001). "Regulation of carbon and electron flow in Clostridium butyricum VPI3266 grown in glucose-glycerol mistures." J. Bacteriology 183: 1748-1754.

111

Silva, E. M.S. T. Yang (1995). "Kinetics and stability of a fibrous-bed bioreactor for continuous production of lactic acid from unsupplemented acid whey." J. Biotechnol. 41: 59-70.

Tanisho, S.;Y. SuzukiN. Wakao (1987). "Fermentative hydrogen evolution by Entebacter aerogenes E. 82005." Int. J. Hydrogen Energy 12(9): 623-627.

Tejavadi, S.M. Cheryan (1988). "Downstream processing of lactic acid whey permeate fermentation broths by hollow fiber ultrafiltration." Appl. Biochem. Biotechnol. 19(1): 61-70.

Wang, F. H.;H. J. Qu;D. W. Zhang;P. TianT. W. Tan (2007). "Produciton of 1,3- propanediol from glycerol by recombinant E.coli using incompatible plasmids system." Mol. Biotech. 37(2): 112-119.

Wu, Z., Yang, S.T. (2003). "Extractive fermentation for butyric acid production from glucose by Clostridium tyrobutyricum." Biotech. Bioeng. 82: 93-102.

Wu, Z. T.S. T. Yang (2003). "Extractive fermentation for butyric acid production from glucose by Clostridium tyrobutyricum." Biotech. Bioeng. 82: 93-102.

Yang, S. T. (1996). Extractive fermentation using convoluted fibrous bed bioreactor. U.S. Patent NO. 5563069.

Yang, S. T.;Y. HuangG. Hong (1995). "A novel recycle batch immobilized cell bioreactor for propionate production from whey lactose." Biotech. Bioeng. 45: 379-386.

Yang, S. T.;H. Zhu;Y. LiG. Hong (1994). "Continuous propionate production from whey permeate using a novel fibrous bed bioreactor." Biotech. Bioeng. 43(11): 1124-1130.

Zhang, T.;H. LiuH. H. P. Fang (2003). "Biohydrogen production from starch in wastewater under thermophilic conditions. ." J. Environ. Manage. 69: 149-156.

Zhu, Y.;X. LiuS. T. Yang (2004). "Construction and characterization of pta gene-deleted mutant of Clostridium tyrobutyricum for enhanced butyric acid fermentation." Biotech. Bioeng. 90: 154-166.

Zhu, Y.;Z. T. WuY. S.T. (2002). "Butyric acid production from acid hydrolysate of corn fiber by Clostridium tyrobutyricum in a fibrous-bed bioreactor." Proc. Biochem. 38(5): 657-666.

Zhu, Y.S. T. Yang (2003). "Adaptation of Clostridium tyrobutylicum for enhanced tolerance to butyric acid in a fibrous-bed bioreactor." Biotech. Prog. 19: 365-372.

112

Zigova, J.;E. Sturdisk;D. VandakS. Schlosser (1999). "Butyric acid production by Clostridium butyricum with integrated extraction and pertraction." Proc. Biochem. 34: 153-160.

113

Vitamin Fe Ni Metal Yeast Trypticase Trypton Butyric acid Gas Volume Growth rate

(Yes or No) (Yes or No) (Yes or No) (Yes or No) (0 – 5 g/l) (0 – 5 g/l) (0 – 5 g/l) (g/l) (ml) (h-1) 1 N N N N 5 5 5 0.177 28 0.138 2 Y Y Y Y 0 0 0 0.057 3 0 3 Y Y N Y 5 5 0 0.168 24 0.147 4 N Y Y Y 0 5 5 0.107 11 0.124 5 Y N N N 5 0 0 0.123 12 0.098 6 N N N N 5 5 5 0.158 29 0.14 7 N N Y Y 5 5 0 0.169 30 0.145 8 Y Y N Y 5 5 0 0.161 24 0.136 9 Y Y N N 0 0 5 0.08 2 0 10 N N Y N 0 0 5 0.077 2 0 11 Y N Y Y 5 0 5 0.122 21.5 0.106 12 Y Y Y Y 2.5 2.5 2.5 0.122 25 0.132 13 Y Y Y Y 0 0 0 0.098 2.5 0 14 Y Y Y Y 2.5 2.5 2.5 0.144 17 0.132 15 N Y N N 0 5 0 0.099 9 0.135 16 N Y Y N 5 0 0 0.132 15 0.096 17 N Y N N 0 5 0 0.123 10 0.136 18 N N Y N 0 0 5 0.099 3 0 19 N N Y Y 5 5 0 0.206 29.5 0.118 20 N Y N Y 5 0 5 0.167 22 0.123 21 Y N N N 5 0 0 0.139 11 0.081 22 N Y N Y 5 0 5 0.108 2 0.005 23 Y N N Y 0 5 5 0.139 14.5 0.142 24 N Y Y Y 0 5 5 0.149 17 0.153 25 Y Y Y N 5 5 5 0.354 33 0.129 26 N N N Y 0 0 0 0.135 2 0 27 N Y Y N 5 0 0 0.193 18 0.112 28 Y Y Y N 5 5 5 0.266 33 0.132 29 Y Y Y N 2.5 2.5 2.5 0.196 14.5 0.134 30 Y N Y N 0 5 0 0.172 8 0.137 31 Y N N Y 0 5 5 0.173 12 0.107 32 N N N Y 0 0 0 0.158 2 0 33 Y Y N N 0 0 5 0.086 1 0 34 Y Y Y N 2.5 2.5 2.5 0.158 17.5 0.17 35 Y N Y N 0 5 0 0.105 2 0 36 Y N Y Y 5 0 5 0.147 21.5 0.112

Table 4.1 Experimental Design and responses for screening medium components to simplify the medium composition (Y: Yes; N: No. Units for the factors and the responses: yeast, g/l; soy broth, g/l; trypton, g/l; butyric acid, g/l; hydrogen, ml; growth rate, /h.

114

Strain Wild type ptb mutant

Substrate Glucose Xylose Glucose Xylose

Product yield (mol/mol)

Butyrate 0.74±0.03 0.61±0.07 0.74±0.01 0.59±0.04

Acetate 0.13±0.02 0.18±0.04 0.20±0.03 0.24±0.06

Hydrogen 1.43±0.17 1.06±0.06 1.60±0.04 1.17±0.06

CO2 1.12±0.13 1.12±0.10 1.00±0.03 1.15±0.03

Butyrate/Acetate (mol/mol) 2.95±0.31 3.44±0.33 1.93±0.29 2.47±0.41

Productivity (mmol/l·h)

Butyrate 1.80±0.67 1.66±0.63 2.27±0.43 2.45±0.67

Acetate 0.74±0.28 0.43±0.31 1.34±0.01 0.88±0.06

Hydrogen 5.15±1.37 4.26±1.06 7.12±2.21 6.23±0.2

Specific growth rate (h-1) 0.15±0.02 0.10±0.02 0.20±0.04 0.14±0.01

All data are based on the average from duplicate batch fermentations with the standard error indicated by the number after ±.

Table 4.2 Comparison of product yields, productivities, and specific growth rates of wild type and ptb mutant in batch fermentations with glucose and xylose

115

ptb mutant free cell ptb mutant FBB fed- fed batch batch overall overall

butyrate yield (mol/mol glucose) 0.59 ± 0.04 0.66 ± 0.02

acetate yield (mol/mol glucose) 0.12 ± 0.01 0.18 ± 0.01

hydrogen yield (mol/mol glucose) 1.68 ± 0.07 1.90 ± 0.10

butyrate concentration (g/l) 21.3 ± 1.3 31.5 ± 2.2

acetate concentration (g/l) 5.4 ± 0.3 11.0 ± 0.5

B/A (mol) 2.7 ± 0.1 1.95 ± 0.05 All data are based on the average from duplicate batch fermentations with the standard error indicated by the number after ±.

Table 4.3 Comparison of product yields, final product concentration, B/A ratio from free-cell and FBB fermentations by ptb mutant.

116

A

B

Figure 4.1 Construction of fibrous-bed bioreactor (A. Schematic drawing; B. Photograph)

117

Figure 4.2 Results of factor screening by design of experiments.

118

1.8

1.6 CGM ) 1 ‐ 1.4 SCGM (h µ 1.2 or

1 0.8 0.6 (mol/mol)

0.4 Yield 0.2 0 Specific growth rate Butyric acid yield Hydrogen yield

Figure 4.3 Comparison of the specific growth rate, butyric acid yield, and hydrogen yield from the simplified medium and the original CGM medium.

119

6 O.D. 5 acetic acid /O.D. 4 butyric acid (g/l)

3

2 Concentration 1

0 basal soluble cellulose glucose xylose lactose glycerol medium starch

Figure 4.4 Cell growth and organic acid production from different substrates.

120

2.5 A 2

1.5 10 g/l 600 30 g/l OD 1 50 g/l

0.5 70 g/l 90 g/l 0 0 10203040 Time (h)

0.25 ) 1 ‐ B (h 0.2 µ

rate 0.15

0.1 growth

0.05 Specific 0 10 30 50 70 90

Glucose concentration (g/l)

Figure 4.5 Effect of the initial glucose concentration on cell growth in RCM. (A) The growth curve; (B) The specific growth rate in the exponential phase.

121

2.5 pH 4 A 2 pH 5 pH 6 1.5 pH 7

600 1 pH 8

OD pH 9 0.5

0

‐0.5 0 10203040 Time (h)

0.25 B ) 1 ‐ 0.2 (h

rate 0.15

growth 0.1

0.05 Specific

0 456789 pH

Figure 4.6 Effect of pH on cell growth: (A) The growth curve; (B) The specific growth rate in the exponential phase.

122

35 Wild‐type with glucose 6000

30 5000

25 4000 20 3000 15 2000

Concentration (g/L) 10 Gas production (mL)

5 1000

0 0 0 20406080100 Time (h)

50 ptb with glucose 10000

40 8000

30 6000

20 4000 Concentration (g/L) Gas production (mL) 10 2000

0 0 0 20406080100 Time (h) O.D. glucose acetate butyrate H2 CO2

Figure 4.7 Fermentation kinetics of glucose by wild type and ptb mutant in free-cell fermentation.

123

60 Wild‐type with xylose 7000

50 6000 5000 40 4000 30 3000 20 2000 Concentration (g/L) Concentration Gas production (mL) production Gas 10 1000

0 0 0 306090120150 Time (h)

60 ptb with xylose 10000 9000 50 8000 7000 40 6000 30 5000 4000 20 3000 Gas production (mL) production Gas Concentration (g/L) Concentration 2000 10 1000 0 0 0 306090120150 Time (h)

O.D. xylose acetate butyrate H2 volumn CO2

Figure 4.8 Fermentation kinetics of xylose by wild type and ptb mutant in free cell fermentation

124

60 16000 A 600 14000 50 OD

12000 (ml)

or 40 10000 (g/l)

30 8000 volume

6000 20 4000

10 Hydrogen 2000 Concemtration 0 0 0 50 100 150 200 250 Time (h)

OD glucose aceticacid butyric acid hydrogen

25 16000 14000 (g/l) 20 (ml)

12000

15 10000

8000 volume

10 6000 concentration

4000 5 2000 Hydrogen Product 0 0 0 1020304050607080 Glocose consumed (g/l)

butyric acid aceticacid hydrogen

Figure 4.9 Fermentation kinetics of fed-batch free-cell fermentation of glucose by the ptb mutant at 37°C, pH 6.0.

125

60 50000 45000 50 40000 35000 40 30000 30 25000 20000 20 15000 Concentration (g/l)

10000 Hydrogen volume (ml) 10 5000 0 0 0 20 40 60 80 100 120 140 160 180 Time (h)

glucose acetic acid butyric acid hydrogen

35 50000 45000 30 40000 25 35000

20 30000 25000 15 20000 10 15000

Concentration (g/l) 10000

5 Hydrogen volume (ml) 5000 0 0 0 20406080100

Glucose consumed (g/l) butyric acid acetic acid hydrogen

Figure 4.10 Fermentation kinetics of fed-batch fermentation by ptb mutant from glucose in the FBB at 37°C, pH 6.0.

126

CHAPTER 5

GENOME SEQUENCING AND METABOLIC ENGINEERING OF

CLOSTRIDIUM TYROBUTYRICUM FOR BUTANOL PRODUCTION

Summary

The genome of Clostridium tyrobutyricum was sequenced using the 454 technology. Genes encoding enzymes involved in acid and solvent formation pathways were annotated. The only missing gene for a complete butanol biosynthesis route is butyrylaldehyde dehydrogenase. Moreover, the mutants of C. tyrobutyricum showed higher tolerance to butanol than the butanol-producing Clostridium acetobutylicum. Two expression plasmids carrying aldehyde/alcohol dehydrogenase gene aad and adhE2, respectively, from C. acetobutylicum ATCC 824 were constructed and introduced into C. tyrobutyricum. The engineered mutants showed significant butanol production in batch fermentation with glucose as the substrate. This work demonstrated that it is feasible to use C. tyrobutyricum to produce butanol.

5.1 Introduction

Recently, the crude oil price has increased and fluctuated dramatically.

Consequently, biobutanol and other biofuels have become more attractive to the fuel industry. Butanol is regarded as a better fuel candidate than ethanol due to the following

127

advantages: (a) butanol has 25% more Btu per gallon; (b) butanol is less evaporative with

a lower vapor pressure; (c) butanol is safer with a higher flash point; (d) butanol has a

higher octane rating; (e) butanol is more miscible with gasoline and diesel but less miscible with water; (f) butanol can be dispersed through existing pipelines and filling stations and used as fuel without any modification to the modern automobile

(Shapovalova and Ashkinazib 2008). In addition, butanol has many industrial applications as solvent, rubber monomer, break fluid, and extractant in food and cosmetics industry (Wackett 2008). Currently, butanol is almost exclusively produced via petrochemical routes and its current price as an industrial solvent is around $ 5.0 per gallon with a huge worldwide market of 1.4 billion gallons per year

(http://www.icispricing.com/il_shared/Samples/SubPage153.asp). The market demand is expected to increase dramatically if butanol can be produced economically to be used as a gasoline substitute.

As an alternative to petroleum-based production, acetone-butanol-ethanol fermentation (ABE fermentation) by the strict anaerobic bacterium C. acetobutylicum was once (1917-1955) one of the largest fermentation processes ever developed in industry (Jones and Woods 1986). In a typical ABE fermentation, butyric and acetic acids are produced first by C. acetobutylicum; the culture then undergoes a metabolic shift and solvents (butanol, acetone, and ethanol) are formed. However, the actual fermentation is quite complicated and suffers from severe final product inhibition, resulting in low reactor productivity, yield and final butanol concentration. Consequently, biobutanol

128

produced from clostridia fermentation is uncompetitive to bioethanol and gasoline in the

fuel market.

The product inhibition is the key issue for butanol fermentation. It is believed

that the saturated/unsaturated fatty acid ratio of the cell membrane has significant impacts

on the solvent tolerance (Bhupinder et al. 1987). The physiological response of bacteria

to the environment with a high butanol concentration is to increase its

saturated/unsaturated fatty acid ratio in order to maintain the rigidity of cell membrane

(Lenaz et al. 1975; Moreira et al. 1981). It has been reported that S. cereviciae with a higher ethanol tolerance also had a higher saturated/unsaturated fatty acid ratio in the cell membrane (Hilgc-Rotmann and Rchm 1991). A similar finding was reported for C. tyrobutyricum, which acquired a higher saturated/unsaturated fatty acid ratio and became more tolerable to butyric acid after adaptation in the fibrous bed bioreactor (FBB) (Zhu et al. 2002). Furthermore, several C. tyrobutyricum mutants with disrupted acetate pathway also showed a higher tolerance to butyric acid (Zhu et al. 2004; Liu and Yang 2006) and butanol (in this study).

The genome of C. tyrobutyricum has been sequenced using the 454 technology.

Some genes encoding the enzymes involved in solventogenesis were found to also exist in C. tyrobutyricum genome (data not published), although C. tyrobutyricum does not produce solvent. Those genes annotated in the 454 database include alcohol dehydrogenase, butanol dehydrogenase, and CoA transferase. In order to construct a complete butanol biosynthesis pathway in C. tyrobutyricum, aldehyde dehydrogenase needs to be expressed and functional in the bacterium. It is thus hypothesized that the

129 introduction of the aad gene encoding aldehyde/alcohol dehydrogenase into C. tyrobutyricum could yield a mutant producing butanol. Furthermore, partial blockage of the acetate formation pathway in pta and/or ack mutant could lead more carbon source to butyrylCoA and butanol production.

In this work, the 454 sequencing results were analyzed and the butanol production pathway was constructed by introducing either aad or adhE2 gene of C. acetobutylicum

ATCC 824. Two expression plasmids were constructed: pCAAD containing aad gene and its native promoter and pSOS-adhE2 containing adhE2 gene and a constitutive ptb promoter. After introducing these plasmids into C. tyrobutyricum by electroporation, several mutants were selected and tested for butanol production in batch fermentation.

The positive result from the fermentation kinetics study confirmed that introducing aad or adhE2 into C. tyrobutyricum enabled the mutant to produce butanol from glucose.

5.2 Materials and Methods

5.2.1 Strains, plasmids and media

C. tyrobutyricum ATCC 25755 was obtained from the American Type Culture

Collection. The plasmids pCAAD and pSOS94 were obtained from Dr. Papoutsakis’ group at the University of Delaware (Nair and Papoutsakis 1994; Sillers et al. 2009).

These two plasmids contained both gram-negative replicator ColE1 and gram-positive replicator RepL, although not in the same arrangement, so they can be replicated in both

E. coli and C. tyrobutyricum. They also carried two selective markers, erythromycin 130

resistant gene for C. tyrobutyricum and ampicillin resistant gene for E. coli (Table 5.1).

The pSOS94 also contains the acetone operon including ctfA+ctfB+adc structure gene.

C. tyrobutyricum was maintained in RCM (Reinforced Clostridium Medium,

Difco) under anaerobic conditions at 4°C. The mutant colonies were screened on RCM

agar plate supplemented with 40 µg/ml erythromycin at 37°C in an anaerobic chamber.

Unless otherwise noted, RCM supplemented with glucose as the carbon source was used

in all studies. The P2 basal medium contains (per 100 ml): 0.5 g of K2HPO4, 0.5 g of

KH2PO4, 2.2 g of CH3COONH4, 2.0 g of MgSO4·7H20, 0.1 g of MnSO4· H20, 0.1 g of

NaCl, 0.1 g FeSO4·7H20, 100 mg of p-aminobenzoic acid, 100 mg of thiamine and 1 mg

of biotin (Mermelstein and Papoutsakis 1993) was also used in fermentation kinetics study.

5.2.2 Genome DNA extraction

Chromosome DNA of C. tyrobutyricum was isolated using the QIAGEN genomic

DNA kit (Qiagen, Valencia, CA). The DNA/protein ratio was checked by measuring O.D.

260/280. Agarose gel was also used to check the integrity of the chromosome DNA

(Figure 5.2). Plasmid DNA in E. coli was isolated using QIAprep Miniprep plasmid purification kit (Qiagen, Valencia, CA) and used for transformation purposes.

5.2.3 454 sequencing and metabolic pathway reconstruction

131

454 sequencing was performed in Craig Venter Institute (Rockville, MD). The

detailed sequencing process could be found in 454’s website

(http://www.454.com/products-solutions/how-it-works/index.asp).

Genes encoding the enzymes involved in the metabolic pathway were annotated

in 454 databases and Pairwise Sequence Alignment was performed between C.

tyrobutyricum and C. acetobutylicum ATCC 824 (NC003030 & NC001988).

5.2.4 pCAAD and pSOS-adhE2 construction

The plasmid pCAAD (Figure 5.3) was used as received. It was constructed by

inserting the aad gene with its own putative promoter into plasmid pIM13. The plasmid

pSOS-adhE2 (Figure 5.4) was constructed by inserting adhE2 gene into plasmid pSOS95

with the constitutive ptb promoter isolated from C. acetobutylicum ATCC 824.

The adhE2 gene was amplified by PCR using primers PE1 and PE2 by pfx

polymerase under standard conditions. The PCR amplification product and plasmid

pSOS94 were digested with restriction enzymes BamHI and SfoI and ligated with T4 to generate recombinant pSOS-adhE2.

PE1(forward): 5’- ATGGATCCTTTTATAAAGGAGTGTATATAAATGAAAG -3’ BamH I RBS PE2-reverse: 5’- TTGGCGCCATAATGAAGCAAAGACTATTTTACATTC -3’ Sfo I

132

The recombinant plasmid pSOS-adhE2 was transformed into Escherichia coli

DH5α using chemical transformation and selected on Luria–Bertani (LB) plates (1% (w/v)

Trypton, 0.5% (w/v) yeast extract, 0.5% (w/v) NaCl, 2% (w/v) agar) containing 100

μg/ml ampicillin. Transformants were selected and screened by colony PCR using the

above primers. The recombinant plasmids were isolated from positive transformants with

a Maxi-Prep Kit (Qiagen). The gene adhE2 with ptb promoter was sequenced completely, proving that no mutation occurred during the PCR.

5.2.5 Restriction endonuclease system identification

Protoplast and cell crude extract were prepared to detect the restriction endonuclease digestion system in C. tyrobutyricum. ATCC 824, which has active

restriction endonuclease system, was used as positive control (Mermelstein and

Papoutsakis 1993). Protoplast were prepared according to the method previously

described (Zhu 2003). Cell crude extracts were prepared from a 20 ml late-exponential phase culture using sonication. The supernatant was stored at -20 ºC.

For restriction assay using protoplast, 500 ng plasmid pCAAD was mixed with 0

µl, 1 µl, 2 µl, or 5 µl of protoplast and incubated for overnight at 37 ºC in reaction buffer prepared as previously described (Zhu 2003). The result was examined by electrophoresis with 0.8% agarose gel.

For restriction assay using cell crude extract, 500 ng plasmid pSOS-adhE2 was mixed with 2 µl of crude extract (approx. 2 µg protein) and incubated for 18 h at 37 ºC in

133 restriction enzyme buffers 2, 3 and 4 from New England Biolab. The products were analyzed by electrophoresis with 0.8% agarose gel.

5.2.6 Transformation

All manipulations were operated in an anaerobic chamber (Thermo Fisher

Scientific). Cells were grown in 50 ml of RCM supplemented with 40 mM DL-threonine and harvested at OD600=0.6 (Zhu et al. 2004; Jiraskova et al. 2005), then washed with 30 ml of ice-cold electroporation SMP buffer twice and resuspended in 1 ml of SMP buffer in 0.4 cm electroporation cuvette (Bio-Rad). The SMP buffer (pH 7.4) consisted of 270 mM sucrose, 1 mM MgCl2, and 7 mM sodium phosphate. The MgCl2 solution was prepared and autoclaved separately before adding to the sucrose and phosphate solution to prevent precipitation. Two micrograms of ice-cold plasmid pCAAD or pSOS-adhE2 were added into the cuvette containing 1 ml of cell and mixed. The mixture was incubated on ice for 5 min before the pulse applied. The transformation parameters were set as follows: 2.5 kV, 600 Ω, 25 AF. The transformed cells were transferred into pre- warmed RCM, incubated at 37°C for 2 hours, and then plated on RCM plates containing

40 µg/ml erythromycin. Plates were incubated in an anaerobic incubator at 37°C for 3–5 days to develop the mutant colonies.

5.2.7 Butanol and butyrate tolerance

134

Butanol tolerance was evaluated by the inhibition effect of butanol on cell growth

in serum tubes. The experiments were performed on C. tyrobutyricum wild type and

mutants (PTA, PAK, and HydEm). Butanol (0, 3, 6, 9, 12, 15 g/l) was added into 10 ml

of RCM, which was then inoculated with 0.5 ml of freshly prepared C. tyrobutyricum

culture in the exponential phase (OD600=1.0) and maintained at 37°C. Samples were taken at regular intervals to measure the OD600 with a spectrophotometer (Sequoia-Turner, model 340). The specific growth rates at various initial butanol concentrations were determined from the cell density.

Butyrate tolerance was also evaluated by the inhibition effect of butyrate on cell growth in serum tubes. Butyric acid, 0-80 g/l with 10 g/l intervals, was added into 10 ml of RCM and the pH was adjusted to 6. Each tube was then inoculated with 0.5 ml of freshly prepared C. tyrobutyricum culture in the exponential phase (OD600=1.0) and

maintained at 37°C. Samples were taken at regular intervals to measure the OD600 with a spectrophotometer (Sequoia-Turner, model 340). The specific growth rates at various initial butyrate concentrations were determined from the OD600.

5.2.8 Fermentation kinetics

The effect of introducing pCAAD and pSOS-adhE2 into C. tyrobutyricum on

fermentation kinetics was evaluated in serum bottles. Each bottle containing 50 ml of

RCM or P2 medium with 10 g/l glucose was inoculated with 0.5 ml of fresh culture and incubated at 37°C. Samples were taken at regular time intervals and the alcohol content

135

was analyzed with gas chromatography (GC) (Shimadzu, GC2014) as described

previously.

Free-cell fermentation of pCAAD mutant was further studied in a 2-liter stirred

bottle containing 1 L of RCM supplemented with 50 g/l glucose. The medium was sparged with nitrogen for 2 hours to achieve anaerobiosis. The inoculum was 50 ml of fresh cell suspension in the exponential phase. Batch fermentations were carried out at

37ºC and started at pH 6.0, and the pH was maintained at 4.5 adjusted with 6 N NaOH.

Liquid samples were taken at regular intervals to analyze the cell density (OD600) and the concentrations of the substrate and products in the fermentation broth.

5.2.9 Analytical methods

The cell density was calculated based on OD600 measured with a

spectrophotometer (Sequoia-Turner, model 340). One unit of OD600 was corresponding to

0.68 g/l of cell dry weight.

The composition of alcohol produced in the fermentation was analyzed with a gas

chromatograph (Shimadzu GC 2014) equipped with Flame Ionization Detector and

Stabilwax-DA column (Restek) with helium as the carrier gas at the flow rate of 2 ml/min.

The injector temperature and the detector temperature were set to 200°C. The column

temperature was programmed as the followings: 80°C hold for 3 min, and then increased

to 150 °C at 30°C/min and hold at 150°C for 4 min.

136

5.3 Results

5.3.1 Genome sequencing

The 454 sequencing produced 211,099 good reads, containing 55,970,852 bp of

raw sequence. Assembly produced 69 contigs, 3,016,691 bp.

The genes encoding the enzymes involved in the metabolic pathway (Figure 5.1)

from pyruvate to various end products, including acidogenesis and solventogenesis

according to ATCC 824, found and annotated in the 454 database are summarized in.

Table 5.2 with the gene name, enzyme name, EC number, contig number, and starting

and ending positions in the 454 database. Among the 16 genes in the pathway, ptb and

buk were missing, which are probably located in the gap due to the disadvantage of the

454 technology. BLAST has been performed against the corresponding gene from

Clostridium. The amino acid sequences of these enzymes display high similarity (60%-

80%) among the clostridium genus compared.

More interestingly, C. tyrobutyricum genome contains some genes necessary for solvent production, such as adc (acetoacetate decarbonase), bdh (NADH dependent butanol dehydrogenase) and adh (NADH dependent alcohol dehydrogenase) (Figure 5.1).

However, the genes encoding alcohol/aldehyde dehydrogenase (aad) cannot be located.

The amino acid sequences from aad and adhE2 are available for C. acetobutylicum

ATCC824. The deduced amino acid sequences from several bifunctional aldehyde alcohol dehydrogenase genes from GenBank database were compared. 66% identity was

found between adhE2 and aad; 57% identity between adhE2 and adhE from E. coli K-12

137

(Fontaine et al. 2002). It was thus proposed that C. tyrobutyricum will probably obtain

the ability to produce butanol if the gene encoding alcohol/aldehyde dehydrogenase is

introduced.

5.3.2 Restriction endonuclease system identification

The restriction enzyme activity has been demonstrated existing in C.

acetobutylicum ATCC 824 with both protoplast and cell crude extract (Mermelstein and

Papoutsakis 1993). However, no digestion of plasmid was observed in the incubation of plasmid pCAAD with the protoplast of C. tyrobutyricum (Figure 5.5A). Furthermore, the crude extract from C. tyrobutyricum did not show any digestion of the plasmid pSOS- adhE2 (Figure 5.5B). These results confirmed that there was no restriction enzyme activity in C. tyrobutyricum.

5.3.3 Transformation

C. tyrobutyricum is a gram-positive bacterium, which is difficult to transform due to its cell wall structure (Allen and Blaschek 1990; Young and Cole 1993).

Electroporation has been demonstrated to be effective to introduce plasmids into gram- positive bacteria (Scott and Rood 1989; Phillips-Jones 1995; Jiraskova et al. 2005).

Electroporation was thus used to transform the recombinant plasmids into C.

tyrobutyricum. The results showed a low transformation efficiency of 2-10 transformants

138 per μg DNA. Nevertheless, several mutant colonies were obtained on the selection agar plates. They were selected and subcultured in CGM containing erythromycin before testing for their butanol production in batch fermentation kinetics study.

5.3.4 Butanol tolerance

All three C. tyrobutyricum mutants (PTA, PAK, HydEm) displayed a higher tolerance to butyrate than the wild type (Figure 5.6). At the same concentration of butyrate, the mutants had a higher specific growth rate and shorter lag phase. As expected, the mutants also showed higher tolerance to butanol than the wild type (Figure 5.7), mutants could tolerate 12 g/l butanol while wild type could not grow at 12 g/l butanol, which is similar to C. acetobutylicum’s butanol tolerance. Figure 5.8 shows all of the three mutants of C. tyrobutyricum shows higher specific growth rate than wild type under various amount of butanol, which supports our hypothesis that the rigid membrane acquired from high butyrate tolerance also contributed to the high solvent tolerance. It is thus reasonable to assume that if we can construct a butanol-producing C. tyrobutyricum mutant, a higher butanol titer probably can be obtained due to its ability to survive at higher butanol concentrations. This higher butanol tolerance was achieved under free-cell fermentation conditions in this study. Based on the experience with butyrate tolerance studies, FBB immobilization could increase the butanol tolerance further.

5.3.5 Fermentation kinetics

139

Butanol and ethanol production in C. tyrobutyricum wild type and two

recombinant heterologous expression strains were studied in RCM and P2 medium in

serum bottles and the results are summarized and compared in Table 5.3. Figure 5.9 shows the time course data on butanol production in batch fermentations by C.

tyrobutyricum wild type and mutant pCAAD in serum bottles. It is clear that the mutant

was able to produce a significant amount of butanol while no detectable butanol was

produced by the wild type under the conditions studied. The final butanol concentration

was 67 mg/l and 24 mg/l in RCM and P2 medium, respectively. Similarly, significant

butanol production was also detected in the fermentations with the mutant pSOS-adhE2,

which gave a final butanol concentration of 9 mg/l and 19 mg/l in RCM and P2 medium,

respectively. It has been noticed that Clostridium beijerinkii produced more butanol in P2

than in RCM (data not published), probably due to the nitrogen deficient condition in the

P2 medium. However, the effect of different media on butanol production is species and

strain dependent, and cannot be generalized without further studies. Although butanol

production was low in these fermentations, both recombinant C. tyrobutyricum strains

constructed in this study showed an ability to produce butanol. It is noted that both

mutants also showed increased ethanol production, as compared to the wild type

fermentation. This indicates that the cloned alcohol/aldehyde dehydrogenase has a broad

substrate specificity and can work with both acetyl-CoA and butyryl-CoA. The relatively

high ethanol background in the wild type fermentation broth was attributed to the ethanol

added to the medium with erythromycin.

140

Free-cell batch fermentations in a pH-controlled fermenter were then performed to further evaluate butanol production in C. tyrobutyricum (pCAAD). The first batch was grown at 37°C in RCM and showed no butanol production. It has been reported that solvent production benefits from slower growth. The temperature was thus decreased to

30°C. Figure 5.10 shows that butanol was produced although at a low level and butyric and acetic acids remained the main fermentation products. The final butanol concentration was 19 mg/ml. Meanwhile ethanol was also produced and reached 60 mg/l.

It’s probably due to the selectivity of AAD on ethanol than butanol, while ADHE2 showed a selectivity on butanol than ethanol (Fontaine et al. 2002). The results confirmed that the cloning of aad into C. tyrobutyricum enabled the mutant to produce both butanol and ethanol, but at a much lower level than those of the native acid products. It is thus necessary to further optimize the fermentation conditions and engineer the mutant to direct more carbon flux toward alcohol synthesis instead of acid production, which can be done by knocking-out pta and in the acid-forming pathways and using a stronger promoter for aad overexpression.

5.4 Discussion

Butanol production from biomass is a promising solution for the energy crisis.

However, the low yield and low productivity due to the tightly regulated metabolic pathway and the final product inhibition make the biobutanol from the ABE fermentation process uncompetitive to gasoline and bioethanol. Many efforts have been made to

141

improve butanol production in ABE fermentation via metabolic engineering of C.

acetobutylicum and process engineering to alleviate inhibition caused by butanol. Despite all these efforts, there has been little progress towards making ABE fermentation economically competitive. A new process for the production of biobutanol is thus needed.

The solvent-producing genes have been cloned, sequenced and characterized in C.

acetobutylicum, including aad, ctfA/B, adc encoding aldehyde/alcohol dehydrogenase,

CoA transferase A/B, acetoacetate decarboxylase, respectively. These genes are located in the megaplasmid pSOL1 in C. acetobutylicum. The aad gene has 2619 base-pair and codes for a 96,517 Dalton protein. The N-terminal section of AAD shows homology to aldehyde dehydrogenase while the C-terminal shows homology to alcohol dehydrogenase.

The entire amino acid sequence of AAD exhibits 56% identity to the trifunctional protein

ADHE from E. coli (Nair et al. 1994). The introduction of aad-containing plasmid into C. acetobutylicum increased the activity of butanol dehydrogenase, acetaldehyde dehydrogenase and butyraldehyde dehydrogenase significantly and also slightly the ethanol dehydrogenase activity in the mutant. The same plasmid has also been transformed into the M5 mutant of C. acetobutylicum, which does not produce solvent

due to the lack of the megaplasmid. However, butanol production resumed upon the

transformation without any detectable acetone and ethanol (Nair and Papoutsakis 1994).

The other gene adhE2 also located on pSOL1 was cloned and characterized by Fontaine

et al. (Fontaine et al. 2002). It has 2,577-bp encoding for a 94.4 kDa protein. The

expression of adhE2 in the DG1 mutant of C. acetobutylicum (BYDH-, AAD-) restored butanol production, and elevated activities of NADH-dependent butyraldehyde and

142

butanol dehydrogenases were observed. The recombinant ADHE2 protein expressed in E. coli also demonstrated NADH dependent butyraldehyde and butanol dehydrogenase activities. It is the first report that one bacterium has two aad genes. Comparing to aad, the adhE2 shows selectivity on butanol over ethanol.

C. tyrobutyricum does not possess the megaplasmid and is not capable of producing butanol, ethanol and acetone. However, the 454 sequencing result revealed that

C. tyrobutyricum genome contains butanol dehydrogenase. It was thus proposed that the introduction of butyraldehyde dehydrogenase could enable butanol production in C. tyrobutyricum. The engineered C. tyrobutyricum would have a relatively simple butanol fermentation pathway without any acetone production due to the lack of acetoacetate decarboxylase. Moreover, its native strong butyric acid production ability can lead more carbon source to butyryl CoA instead of acetate or ethanol. Thus, the engineered mutant is expected to have a more easily controlled metabolic pathway with high butanol production.

The final product inhibition is the key issue for butanol fermentation. The butanol inhibition is related to the cell membrane lipid composition (Weber and De Bont 1996).

A higher saturated/unsaturated fatty acid ratio can better maintain the membrane’s rigidity and prevent the solvent from entering and disrupting the cell membrane, or even

lysing the membrane (Bhupinder et al. 1987). Similarly, the tolerance to organic acid also relates to the lipid composition of cell membrane, in addition to ATPase activity. When the short-chain organic acid is produced, it may pass the cell membrane and enter the intracellular space. A higher saturated/unsaturated fatty acid ratio can make the

143

membrane not permeable to the acid. Once the acid enters the cell, it is disassociated and

the released proton is harmful to the cell because it lowers the intracellular pH from

neutral (O'Sullivan and Condon 1999; Cotter and Hill 2003). ATPase is associated with

the cell membrane and could pump extra proton ions from the intracellular space of the

cell to the outside to maintain a proper intracellular environment at the expense of ATP.

It has been reported that the higher tolerance to butyric acid accompanied a higher saturated fatty acids content (Zhu and Yang 2003). Two gene knockout mutants with disrupted acetate pathway and one FBB adaptation mutant from C. tyrobutyricum were constructed previously in our group (Zhu et al. 2004; Liu et al. 2006) and all of them showed higher butyrate tolerance than the wild type. It is expected that mutants with higher butanol tolerance could be constructed upon the introduction of aad or adhE2 into these acid-tolerant mutants.

In this research, butanol production has been demonstrated in C. tyrobutyricum

mutants, although at a low concentration. The low butanol productivity is probably due to

the promoter compatibility. Further studies should include the alcohol/aldehyde

dehydrogenase assay to detect the enzyme activity level and SDS-PAGE to confirm the

gene expression in the mutant. Meanwhile, as the genome sequence of C. tyrobutyricum

is available, efforts will be put on cloning the native constitutive thiolase promoter to

promote the aad/adhE2 overexpression. Furthermore, mutants with knock-out acetate and butyrate pathways would serve as better hosts for butanol biosynthesis and should be investigated.

144

5.5 Conclusion

The genome of C. tyrobutyricum has been sequenced by using the 454 technology.

Most of genes encoding for the enzymes responsible for the metabolic pathway from

glucose to acetate and butyrate have been identified and localized in the 454 database,

except for ptb and buk. Several enzymes involved in solvent formation also exist in C.

tyrobutyricum. The aldehyde/alcohol dehydrogenase genes aad and adhE2 have been

introduced into the wild type of C. tyrobutyricum. The mutant showed butanol production

at 30°C in free cell fermentation. It is thus concluded that the introduction of aad gene and/or adhE2 gene could complete the metabolic pathway needed for alcohol production and lead to the production of butanol in C. tyrobutyricum.

145

5.6 References

Allen, S. P.H. P. Blaschek (1990). "Factors invovled in the electroporaton-induced transformation of Clostridium perfringens." FEMS Microbiol Letter 58: 217-220.

Bhupinder, K. S.;D. KumudeswarG. Tarun (1987). "Inhibitory factors involved in acetone-butanol fermentation by Clostridium saccharoperbutylacetonicum." Current Microbiol. 16: 61-67.

Cotter, P. D.C. Hill (2003). "Surviving the acid test: response of gram-positive bacteria to low pH." Microb. Molec. Bio. Rev. 67(3): 429-453.

Fontaine, L.;I. Meynial-Salles;L. Girbal;X. H. Yang;C. CrouxP. Soucaille (2002). "Molecular characterization and transcriptional analysis of adhE2, the gene encoding the NADH-dependent aldehyde/alcohol dehydrogenase responsible for butanol production in alcohologenic cultures of Clostridium acetobutylicum ATCC 824." J. Bacteriol 184(3): 821-830.

Hilgc-Rotmann, B.H. J. Rchm (1991). "Relationship between fermentation capability and fatty acid composition of free and immobilized Saccharomyces cerevisiae." Appl. Microbio. Biotech. 34: 502-508.

Jiraskova, A.;L. Vitek;J. Fevery;T. RumlP. Branny (2005). "Rapid protocol for electroporation of Clostridium perfeingens." J. Microbiol Methods 62: 125-127.

Jones, D. T.D. R. Woods (1986). "Acetone-butanol fermentation revisited." Microb. Rev. 50(4): 484-524.

Lenaz, G.;C. G. ParentiA. M. Sechi (1975). "Lipid-protein interactions in mitochondria: changes in mitochondrial adenosine triphosphatase activity induced by n-butyl alcohol." Arch. Biochem. Biophys 167: 72-79.

Liu, X.;Y. ZhuS. T. Yang (2006). "Construction and characterization of ack deleted mutant of Clostridium tyrobutyricum for enhanced butyric acid and hydrogen production." Biotech. Prog. 22: 1265-1275.

Liu, X. G.S. T. Yang (2006). "Kinetics of butyric acid fermentation of glucose and xylose by Clostridium tyrobutyricum wild type and mutant." Proc. Biochem. 41(4): 801-808.

Mermelstein, L. D.E. T. Papoutsakis (1993). "In vivo methylation in Escherichia coli by the Bacillus subtilis phage_3TI methyltransferase to protect plasmids from restriction upon transformation of Clostridium acetobutylicum ATCC 824." Appl. Environ. Microbiol. 59: 1077-1081.

146

Moreira, A. R.;D. C. UlmerJ. C. Linden (1981). "Butanol toxicity in the butyric fermentation." Biotech. Bioeng. Syrup 11: 567-579.

Nair, R. V.;G. N. BennettE. T. Papoutsakis (1994). "Molecular characterization of an aldehyde/alcohol dehydrogenase gene from C. acetobutylicum ATCC 824." Journal of Bacteriology 176: 871.

Nair, R. V.E. T. Papoutsakis (1994). "Expression of plasmid-encoded aad in clostridium M5 restores vigorous butanol production." Journal of Bacteriology 176: 5843.

O'Sullivan, E.S. Condon (1999). "Relationship between acid tolerance, cytoplasmic pH, and ATP and H+-ATPase levels in chemist cultures of Lactococcus lactis." Appl. Environ. Microbiol. 65: 2287-2293.

Phillips-Jones, M. K. (1995). "Introduction of recominant DNA into Clostridium spp. ." Methods Mol. Biol. 47: 227-235.

Scott, P. T.J. I. Rood (1989). "Electroporation-mediated transformation of lysostaphin- treated Clostridium perfringens." Gene 82: 327-333.

Shapovalova, O. I.L. A. Ashkinazib (2008). "Biobutanol: Biofuel of second generation." Russian J. Appl. Chem. 81(12): 2232-2236.

Sillers, R.;M. Al-HinaiE. T. Papoutsakis (2009). "Aldehyde-alcohol dehydrogenase and/or thiolase overexpression coupled with CoA transferase downregulation lead to higher alcohol titers and selectivity in Clostridium acetobutylicum fermentations." Biotechnol. Bioeng. 102(1): 38-49.

Wackett, L. P. (2008). "Microbial-based motor fuels: science and technology." Microb. Biotech. 1(3): 211-225.

Weber, F. J.A. M. De Bont (1996). "Adaptation mechanisms of microorganisms to the toxic effects of organic solvents on membranes." Biocheimica et Biophysica Acta, Reviews on Biomembranes 1286(3): 225-245.

Young, M.S. Cole (1993). Bacillus subtilis and other gram-positive bacteria: biochemistry, physiology, and molecular genetics. Washington D.C., American Society for Microbiology.

Zhu, Y. (2003). Enhanced butyric acid fermentation by Clostridium tyrobutyricum immobilized in a fibrous-bed bioreactor. Department of chemical engineering. Columbus OH, The Ohio State University. Ph.D. dissertation.

147

Zhu, Y.;X. LiuS. T. Yang (2004). "Construction and characterization of pta gene-deleted mutant of Clostridium tyrobutyricum for enhanced butyric acid fermentation." Biotech. Bioeng. 90: 154-166.

Zhu, Y.;Z. T. WuY. S.T. (2002). "Butyric acid production from acid hydrolysate of corn fiber by Clostridium tyrobutyricum in a fibrous-bed bioreactor." Proc. Biochem. 38(5): 657-666.

Zhu, Y.S. T. Yang (2003). "Adaptation of Clostridium tyrobutylicum for enhanced tolerance to butyric acid in a fibrous-bed bioreactor." Biotech. Prog. 19: 365-372.

148

Strains / Plasmids Characteristics Source or reference

Clostridium tyrobutyricum Wild type ATCC ATCC 25755

Clostridium acetobutylicum Wild type ATCC ATCC 824

E. coli DH5α Invitrogen pCAAD Emr, Apr, aad, ColE1, Guerout-Fleury et al. 1995 RepL pSOS94 Emr, Apr, ColE1, RepL, Soucaille and Papoutsakis, ptb promoter, acetone unpublished operon pSOS-adhE2 Emr, Apr, ColE1, RepL, This study adhE2, ptb promoter

Table 5.1 Bacterial strains and plasmids used in this study.

149

454 sequence location Consensus EC to ATCC Enzyme Gene number Contig Start End 824

Pyruvate ferredoxin oxidoreductase pfor 1.2.7.1 33 59205 62750 67.8%

Phosphate acetyltransferase (phosphotransacetylase) pta 2.3.1.8 12 63334 64335 78.7%

Acetate kinase ack 2.7.2.1 12 64388 65584 77.3%

Thiolase (acetyl-CoA acetyltransferase) atoB 2.3.1.9 37 46302 47483 86.0%

3-Hydroxybutyryl-CoA dehydrogenase hbd 1.1.1.157 22 1589 2524 27.8%

Crotonase crt 4.2.1.55 64 1736 2512 74.4%

Butyryl-CoA dehydrogenase bcd 1.3.99.2 55 44 484 72.5%

Butanol dehydrogenase bdhA (NADH dependent) 1.1.1.- 25 76288 77451 67.8%

Butanol dehydrogenase bdh (NADPH dependent) 1.1.1.- 36 17262 18428 77.6%

AcetoacetylCoA: acetate/butyrate CoA transferase ctf 2.8.3.9 64 129 1682 44.0%

AcetoacetylCoA: acetate/butyrate CoA transferase ctf 2.8.3.9 22 21339 22931 42.0%

Ferredoxin hydrogenase hyd 1.12.7.2 4 181223 182194 38.5%

Phosphate butyltransferase (phosphotransbutyrylase) ptb 2.3.1.19 n/a butyrate kinase buk 2.7.2.7 n/a

Alcohol/aldehydes dehydrogenase adhE2 (NADH dependent) 1.1.1.- n/a

Alcohol/aldehydes dehydrogenase aad 1.1.1.- n/a

Acetoacetate decarboxylase adc 4.1.1.4 n/a

Table 5.2 Location of genes and consensus against ATCC824

150

Medium Strain OD600 Ethanol (mg/l) Butanol (mg/l)

RCM Mutant pCAAD 15.4 ± 1.2 240 ± 22 67 ± 8

Mutant pSOS-adhE2 16.8 ± 1.0 190 ± 26 9 ± 1

Wild type 16.2 ±0 .8 210 ± 35 2 ± 2

P2 Mutant pCAAD 3.2 ± 0.4 65 ± 8 24 ± 2

Mutant pSOS-adhE2 2.9 ± 0.2 56 ± 10 19 ± 3

Wild type 3.1 ± 0.5 32 ± 2 1± 1

Table 5.3 Butanol and ethanol production by C. tyrobutyricum wild type and mutants pCAAD and pSOS-adhE2 grown in RCM and P2 medium in serum bottles at 37oC, initial pH 6.

151

ACETOGENESIS SOLVENTOGENESIS LDH Lactate Pyruvate iLDH Fd H PFOR 2 FdH 2 HYD AK PTA ADH AAD Acetate Acetyl-phosphate Acetyl-CoA Acetaldehyde Ethanol

THIL

CoAT CoAT ADC Acetoacetyl-CoA Acetoacetate Acetone

HBD

3-hydroxy-butyryl-CoA

CRO

CoAT Crotonyl-CoA

BCD

BDH BK PTB AAD Butyrate Butyryl-phosphate Butyryl-CoA Butyraldehyde Butanol

Figure 5.1 The enzymes involved in the pathway from pyruvate to the end products, highlighted enzyme presented in 454 database. (PFOR: pyruvate ferredoxin oxidoreductase, HYD: hydrogenase, AAD: alcohol/aldehyde dehydrogenase, ADH: alcohol dehydrogenase, LDH: lactate dehydrogenase, iLDH: NADH dependent lactate dehydrogenase, PTA: phosphotransacetylase, AK: acetate kinase, THIL: thiolase, CoAT: CoA transferase, ADC: acetoacetate decarboxylase, HBD: 3-hydroxybutyryl CoA dehydrogenase, CRO: crotonase, BCD: butyryl-CoA decarboxylase, BDH: butanol dehydrogenase, PTB: phosphotransbutyrylase, BK: butyrate kinase

152

1 2 3

Figure 5.2 The genomic DNA quality check for 454 sequencing (1: 100 bp-12 kb DNA ladded, 2: original genomic DNA, 3: 25 X dilution)

153

ColE1 aad-pro

AmpR

pCAAD aad 7890 bp

MlsR

RepL

Figure 5.3 Structure of plasmid pCAAD. (RepL: gram-positive replicator; ColE1: Gram- negative replicator; ampr: ampicillin resistant gene; mlsr: erythromycin resistant gene; aad: aad structure gene: aad-pro: aad native promoter)

154

Figure 5.4 Construction of plasmid pSOS-adhE2. (RepL: gram-positive replicator; ColE1: Gram-negative replicator; ampr: ampicillin resistant gene; mlsr: erythromycin resistant gene; adhE2: adhE2 structure gene: ptb-Pro: ptb promoter)

155

A

1 2 3 4 5

B

1 2 3 4 5 6 7 8

Figure 5.5 The restriction endonuclease system check with protoplast and cell crude extracts. A. with protoplast (Lane 1, ladder; Lanes 2-5, pCAAD with C. tyrobutyricum protoplast in NEB buffers with different amounts of protoplast); B. with cell extracts (Lane 1, plasmid only; Lanes 2-4, plamid with C. tyrobutyricum cell extract in NEB buffers; Lanes 5-7, plamid with C. acetobutylicum cell extract in NEB buffers; Lane 8, DNA marker).

156

1.2 1.2 1 1 0.8 0.8 0.6 0.6

OD600 0.4 30g/l butyrate

0.4 0 g/l butyrate OD600 0.2 0.2 0 0 0 50 100 150 200 0 50 100 150 200 Time (h) Time (h)

1 0.8 WT 0.6 PTA 60g/l butyrate

OD600 0.4 ACK 0.2 HYDEM 0 0 100 200 300 Time (h)

Figure 5.6 Growth kinetics of Clostridium tyrobutyricum wild type and mutants ACK, PTA, and HydEm under various amounts of butyrate at pH 6.

157

2 2

1.5 1.5

1 1 OD600 0g/l butanol OD600 6g/l butanol 0.5 0.5

0 0 0 50 100 150 200 0 20406080 Time (h) Time (h)

1.4 1.2 1 WT 0.8 12g/l butanol PTA 0.6

OD600 ACK 0.4 HYDEM 0.2 0 0 50 100 150 200 Time (h)

Figure 5.7 Growth kinetics of Clostridium tyrobutyricum wild type and mutants PTA, PAK, and HydEm under various amounts of butanol at pH6

158

120% (%)

100% rate 80% growth 60%

specific 40%

20% Relative

0% 02468101214 Butanol concentration (g/l)

wt pta ack hydem

Figure 5.8 Comparison of specific growth rate under various amount of butanol between

wild type and mutants of C. tyrobutyricum.

159

0.08 6.5

0.07 6 0.06 (g/l)

0.05 5.5 0.04 pH concentration 0.03 5

0.02 Butanol 4.5 0.01

0 4 0 50 100 150 200 Time (h)

WT‐RCM pCAAD‐RCM WT‐P2 pCAAD‐P2 pH

Figure 5.9 Butanol production of C. tyrobutyricum (pCAAD) and wild type in RCM and P2 medium in serum bottles.

160

35 70

30 60 (mg/l) (g/l) 25 50

20 40

15 30 concentration 10 20 O.D./Concentration

5 10 alcohol

0 0 0 50 100 150 200 250 Time (hours)

O.D. glucose acetate butyrate BUTANOL ethanol

Figure 5.10 Free-cell fermentation kinetics of C. tyrobutyricum (pCAAD) at 30°C,

pH4.5.

161

CHAPTER 6

CONCLUSIONS AND RECOMMENDATIONS

6.1 Conclusions

This study explored the role of ptb in the metabolic pathway of Clostridium tyrobutyricum and demonstrated the feasibility of the integration of metabolic engineering and process development for enhanced hydrogen production. The advantages of the fibrous-bed bioreactor were also confirmed. It also verified the potential application of C. tyrobutyricum as a butanol-producing strain. The important results and conclusions obtained in this study are summarized below.

6.1.1 Metabolic engineering

™ Partial ptb gene was amplified by PCR from the chromosome of C. tyrobutyricum

by degenerated primers. The deduced amino acid sequence showed 100% identity

to the PTB from Clostridium acetobutylicum.

™ A non-replicative integrational plasmid containing ptb fragment disrupted by a

erythromycin resistance gene (Emr) was constructed. The ptb-disrupted mutant

was obtained by homologous recombination. The insertion was confirmed by

nested PCR. 162

™ PTB enzyme activity decreased by 76% in the mutant. However, the ptb

inactivation did not turn off the PTB activity completely. Meanwhile, the BK

activity also decreased by 42%, suggesting that the buk is located at the

downstream of the ptb in the same operon. In contrast, PTA activity in the mutant

increased by 7% and AK activity increased by 29%.

™ The specific growth rate of the mutant increased by 17%. The mutant also showed

50% less sensitive to butyric acid inhibition.

6.1.2 Fermentation kinetics

™ The composition of Clostridium Growth Medium was simplified based on statistic

experimental design. Trace metals solution and vitamin solution did not show

significant effects on cell growth and the production of butyric acid and hydrogen

thus were removed from the recipe. NiCl2 and FeSO4 were kept because they are

critical factors for hydrogenase. They were added into the buffer #2 solution and

the buffer was sterilized by microfiltration. Yeast extract and TrypticaseTM soy

broth, were essential for cell growth and also kept.

™ C. tyrobutyricum ptb mutant can ferment xylose and glucose, but cannot uptake

lactose, glycerol, starch, or cellulose.

™ The initial glucose concentration of 50 g/l and pH 6 were determined to be

optimal for the ptb mutant fermentation.

163

™ The ptb mutant showed a lower B/A ratio, higher hydrogen yield, higher specific

growth rate, higher productivity than the wild type in free-cell fermentations. But

it still produced a significant amount of butyric acid, suggesting another pathway

for butyrate formation.

™ Immobilized-cell fermentation in the fibrous-bed bioreactor increased the final

butyric acid concentration from 21.3 g/l to 31.5 g/l. It demonstrated that the FBB

could adapt the cells for an improved tolerance to butyric acid.

6.1.3 Genome sequence

™ The genome sequencing of C. tyrobutyricum by the 454 technology produced

211,099 good reads, containing 55,970,852 bp of raw sequence. Assembly

produced 69 contigs, 3,016,691 bp.

™ ptb and buk were not found in the annotated open reading frames (ORF) in the

454 database. They probably fell into the gaps among various assembled contigs

due to sequencing limitations of the 454 technology.

™ Two genes encoding for acetoacetyl-CoA acetate/butyrate CoA transferase were

identified, suggesting the existence of CoA route in C. tyrobutyricum and

explaining why the butyrate was still produced in the ptb-disrupted mutant.

™ Several genes involved in the clostridia solventogenesis pathway, such as adc

(acetoacetate decarboxylase), bdh (NADH dependent butanol dehydrogenase) and

adh (NADH dependent alcohol dehydrogenase), also exist in C. tyrobutyricum,

164

suggesting a possible butanol biosynthesis pathway that could be constructed with

the cloning of an alcohol/aldehyde dehydrogenase gene in C. tyrobutyricum.

6.1.4 Butanol production by C. tyrobutyricum

™ C. tyrobutyricum mutants pta, ack, hydem showed a higher butanol tolerance than

the wild type and C. acetobutylicum. The tolerance to butanol and butyrate are

strongly correlated.

™ Two expression plasmids pCAAD and pSOS-adhE2 were constructed. pCAAD

carried aad gene with its native promoter, and pSOS-adhE2 carried adhE2 gene

promoted by the constitutive ptb promoter of C. acetobutylicum. Both plasmids

were transformed into C. tyrobutyricum and significant butanol production was

verified in batch fermentations. This work demonstrated the potential application

of C. tyrobutyricum as a butanol producing bacterium.

6.2 Recommendations

Several attempts have been explored to understand the metabolic pathway of C. tyrobutyricum and to improve butyric acid and hydrogen production. However, many problems remain unsolved and the exact metabolic pathway and its regulation system remain unknown. This section lists several areas suggested for future studies.

6.2.1 Construction of novel butanol producers

165

Transformation of pCAAD and pSOS-adhe2 established the butanol production in

C. tyrobutyricum. However, the butanol concentration in the fermentation broth was quite low, probably due to the low expression efficiency. Thus, the following studies are recommended:

™ Construct a reporter gene plasmid to detect if ptb promoter from C.

acetobutylicum works in C. tyrobutyricum efficiently.

™ Thiolase is constitutively expressed in Clostridia. The thiolase structure gene is

located in contig 37 from 46302 bp to 47483 bp and its upstream nucleotide

sequence is also available in the C. tyrobutyricum 454 database. The suggested

strategy is to clone the native thiolase promoter region and use it for the aad

expression.

™ Increase the copy number of plasmid by changing the current replicator repL from

Bacillus to a native clostridia replicator. Several plasmids with replicator from

Clostridium are available. The expression plasmid could be constructed based on

those plasmids.

™ Transformation is a key technique for the further study of genetic engineering on

C. tyrobutyricum. The current cloning work is limited by the low transformation

efficiency, which is only 10 colonies per microgram plasmid or even lower.

Optimization study suggested that the limiting factor is not the cell viability after

the electrical pulse shock, but the impermeability of the cell wall and cell

membrane. More harsh conditions, such as higher voltage, pulse frequency, or

166

incubation with lysozyme, should be explored to improve the transformation

efficiency.

™ The butanol pathway needs to be established in the mutant (PTA, PAK, HydEm,

PTB) of C. tyrobutyricum in order to acquire higher butanol tolerance.

™ There are only two antibiotics (erythromycin and thiamphenicol) to which C.

tyrobutyricum is known to be sensitive. For the purpose of constructing mutants

with dual or multiple mutations, a selection marker rescue system needs to be

developed in order to reuse the same antibiotic resistance gene repeatedly.

6.2.2 CoA transferase

It was noticed that in the ptb mutant fermentation, the butyrate was still produced

at a comparable level to the wild type fermentation, suggesting the existence of a novel

pathway other than PTB-BK route that could also produce butyrate. The genome of C.

tyrobutyricum has recently been sequenced with the 454 technology and the results

showed the existence of two genes for acetoacetyl-CoA: acetate/butyrate CoA transferase

(EC# 2.8.3.9), one of them is located in the contig 64 starting at 129 bp and ending at

1682 bp and the other is from 21339 bp to 22931 bp in the contig 22. The deduced amino

acid sequences from these two genes show 57% identity. The one located on contig 64

shows 80% identity to the CoA transferase from Clostridium Novyi (Gi: 4541513) and

the other shows 76% identity to the same protein. The function of this enzyme is to

transfer CoA group among butyryl-CoA, acetyl-CoA and acetoacetyl-CoA to form

butyrate, acetate, and acetoacetate depending on the substrate availability. This pathway

167 has been identified as a major butyrate production pathway in several other species such as Roseburia sp. A2-183 (Charrier et al. 2006). It is possible that C. tyrobutyricum possesses both pathways for butyrate biosynthesis, which can explain why ptb-disruption cannot eliminate butyrate production and pta/ack knock out also cannot eliminate acetate production (Zhu et al. 2004; Liu et al. 2006). Further studies of the CoA transferase function in C. tyrobutyricum are thus recommended.

™ Clone and sequence the two CoA transferases from the chromosome of C.

tyrobutyricum.

™ Overexpression of CoA transferase gene in E. coli and evaluate its expression and

function by sodium dodecyl sulfate polyacrylamide gel electrophoresis (SDS-

PAGE), enzyme assay and fermentation kinetics.

™ Construct mutants with overexpression or knock-out of CoA transferase to

evaluate its effects on butyrate and acetate production.

6.2.3 Functional genomics and proteomics

DNA microarray or gene chips have been widely used for the functional analysis of the genome. The whole transcriptional response to gene mutations or environmental stimulus could be examined. This technique can offer the information of the network change compared to the traditional single gene mutation approach.

A shot-gun genome library of C. tyrobutyricum was constructed where the whole genome is represented by random DNA fragments. The library contains 1860 gene

168 fragments with size ranging from 1 kb to 5 kb. The probability of the gene of interest presenting in the library is 90%. Microarray chips have been fabricated onto the poly- lysine coated glass slide and the method for the target preparation and direct labeling has been validated (Appendix A). The microarray is ready to use for studying the global gene expression changes in the following situations:

™ Transcriptome change in the 4 mutants of C. tyrobutyricum. Three different

genetically engineered mutants, including PTA, PAK, PTB, were constructed and

one adaptation mutant HydEm was obtained from the FBB fermentation. For the

three single gene knock-out mutants, the transcriptome change will not be limited

in the gene interrupted, some global change will happen due to the complicated

regulation mechanism in C. tyrobutyricum. For the adaptation mutant, it is

hypothesized that there were several genes mutated in the adaptation process,

such as the H+-ATPase or genes responsible for the cell membrane composition

and lipid biosynthesis. The transcriptome of each mutant and the wild type will be

hybridized against microarray chip competitively and the fragments showing

different expression levels will be identified and sequenced. The entire gene

sequence will be obtained by BLASTing against the 454 database.

™ Transcriptome change in different environments, such as pH, hydrogen partial

pressure, and production concentration. All of these conditions have shown

significantly impacts on the fermentation kinetics so they should have also

induced global gene expression changes. One standard condition will be chosen as

the baseline. The transcriptome from the target condition and the standard

169

condition will be competitively hybridized against microarray chip. The entire

genes sequence will be obtained from the 454 database.

The purpose of the functional genomics analysis by microarray is to understand the metabolic pathway and the gene expression regulation. The results can be used to elucidate the function of key genes and to identify critical genes in the metabolic pathway as targets for further rational metabolic engineering to construct knock-out or overexpression mutants with higher yield, productivity and tolerance.

It’s also recommended to use two-dimensional electrophoresis to confirm the protein expression pattern change among the mutants and under various conditions. The results could support regulation mechanism obtained from functional genomics analysis.

A precise understanding of metabolic pathway will facilitate the further manipulation of

C. tyrobutyricum to improve butyric acid and hydrogen production.

170

6.3 Reference

Charrier, C.;G. J. Duncan;M. D. Reid;G. J. Rucklidge;D. Henderson;P. Young;V. J. Russell;H. F. AminovP. Louis (2006). "A novel class of CoA-transferase involved in short-chain fatty acid metabolism in butyrate-producing human colonic bacteria." Microbiology 152: 179-185.

Liu, X.;Y. ZhuS. T. Yang (2006). "Construction and characterization of ack deleted mutant of Clostridium tyrobutyricum for enhanced butyric acid and hydrogen production." Biotech. Prog. 22: 1265-1275.

Zhu, Y.;X. LiuS. T. Yang (2004). "Construction and characterization of pta gene-deleted mutant of Clostridium tyrobutyricum for enhanced butyric acid fermentation." Biotech. Bioeng. 90: 154-166.

171

BIBLIOGRAPHY

Agreda, V. H.J. R. Zoeller (1993). Acetic acid and its derivatives. New York, M. Dekker.

Alam, S.;D. StevensR. Bajpai (1988). "Production of butyric acid by batch fermentation of cheese whey with Clostridium beijerinckii." J. Indus. Microb. 2(6): 359-364.

Alasfour, F. N. (1997). "Butanol-A single cylinder engine study: engine performance." Int. J. Hydrog. Energy 21: 21-30.

Allen, S. P.H. P. Blaschek (1990). "Factors invovled in the electroporaton-induced transformation of Clostridium perfringens." FEMS Microbiol Letter 58: 217-220.

Alsaker, K.E. T. Papoutsakis (2004). Microarray-based analysis of the stress response in Clostridium acetobutylicum cultures. 227th ACS National Meeting. Anaheim, CA.

Andersch, W.;H. BahlG. Gottschalk (1983). "Level of enzymes involved in acetate, butyrate, acetone and butanol formation in Clostridium acetobutylicum." European Journal of Applied Microbiology and Biotechnology 18(6): 327-332.

Aristidou, A.M. I. Penttila (2000). "Metabolic engineering applicationto renewable resource utilization." Curr. Opin. Biotech. 11: 187-198.

Atsumi, S.;A. F. Canna;M. R. Connora;C. R. Shena;K. M. Smitha;M. P. Brynildsena;K. J. Y. Choua;T. HanaiJ. C. Liao (2008). "Metabolic engineering of Escherichia coli for 1-butanol production." Metabo. Eng. 10(6): 305-311.

Baer, S. H.;H. P. BlaschekT. L. Smith (1987). "Effect of butanol challenge and temperature on lipid composition and membrane fluidity of butanol-tolerant Clostridium acetobutylicum." Appl. Environ. Microbiol. 53(12): 2854-2861.

Bai, M. D.;S. S. ChengY. C. Chao (2004). "Effects of substrate components on hydrogen fermentation of multiple substrates." Water Sci. Tech. 50(8): 209-216.

Bhupinder, K. S.;D. KumudeswarG. Tarun (1987). "Inhibitory factors involved in acetone-butanol fermentation by Clostridium saccharoperbutylacetonicum." Current Microbiol. 16: 61-67.

172

Biebl, H. (2001). "Fermentation of glycerol by Clostridium pasteurianum-batch and continuous culture studies." J. Indus. Microb. & Biotech. 27(1): 18-26.

Bisaillon, A.;J. TurcotP. C. Hallenbeck (2006). "The effect of nutrient limitation on hydrogen production by batch cultures of Escherichia coli." Int. J. Hydrog. Energy 31: 1504-1508.

Bisson, L. F. (1988). "High affinity glucose transport in Saccharomyces cerevisiae is under general glucose repression control." J. Bacteriol 170: 4838-4835.

Bisson, L. F.;D. M. Coons;A. L. KruckebergD. A. Lewis (1993). "Yeast sugar transporters." Crit. Rev. Biochem. Mol. Biol. 28: 259-308.

Bogovic, M. B.;M. Koman Rajsp;B. PerkoI. Rogelj (2007). "Inhibition of Clostridium tyrobutyricum in cheese by Lactobacillus gasseri." Int. Dairy J. 17(2): 157-166.

Book, C.-W. F. The world fact book. CIA, https://www.cia.gov/cia/publications/factbook/rankorder/2174rank.html.

Boyaval, P.;C. CorreS. Terre (1987). "Continuous lactic acid fermentation with concentrated product recovery by ultrafiltration and electrodialysis." Biotech. Lett. 9(3): 207-212.

Boyles, D. (1984). Bioenergy technology thermodynamics and costs. New York, Wiley.

Boynton, Z.;G. BennettR. Rudolph (1996). "Cloning, sequencing, and expressing of genes encoding phosphotransacetylase and acetate kinase from C. acetobutylicum ATCC824." Appl. and Enviro. Microbiol. 62: 2758-2766.

Bro, C.;B. Regenberg;J. ForsterJ. Nielsen (2006). "In silico aided metabolic engineering of Saccharomyces cerevisiae for improved bioethanol production." Metabo. Eng. 8: 102-111.

Bro, C.;B. RegenbergJ. Nielsen (2004). "Genome-wide transcriptional response of a Saccharomyces cerevisiae strain with an altered redox metabolism." Biotechnology and Bioengineering 85: 269-276.

Broder, J. D.;J. W. Barrier;K. P. LeeM. M. Bulls (1995). "Biofuels system economics." World Resour. Rev. 7(4): 560-569.

Brosseau, J. D.J. E. Zajic (1982). "Continuous microbial-production of hydrogen gas." Int. J. Hydrog. Energy 7(8): 623-628.

Bruggemann, H.;S. Baumer;W. F. Fricke;A. Wiezer;H. O. Liesegang;I. Decker;C.

173

Herzberg;R. Martinez-Arias;R. Merkl;A. HenneG. Gottschalk (2003). "The genome sequence of Clostridium tetani, the causative agent of tetanus disease." PNAS 100(3): 1316-1321.

Buelter, T.;A. Hawkins;K. Kersh;P. Meinhold;M. PetersE. Subbian (2008). "Engineered microorganisms for producing n-butanol from glycerol." PCT Int. Appl.: 105.

Cammack, R.;M. FreyR. Robson (2001). Hydrogen as a fuel: Learning from future. London, Taylor & Francis.

Cary, J. W.;D. J. Perterson;E. T. PapoutsakisG. Bennett, N. (1988). "Cloning and expression of Clostridium acetobutylicum phosphotransbutyrylase and butyrate kinase genes in Escherichia coli." J. Bacterial 170: 4613-4618.

Casey, G. P.W. E. Ingledew (1986). "Ethanol tolerance in yeasts." Crit. Rev. Microbiol 13: 219-280.

Cato, E.;W. GeorgeS. Finegold (1986). Genus clostridia prazmowski 1880. Bergey's manual of systematic bacteriology. Baltimore, William & Wilkins Co. 2: 1141-1200.

Charrier, C.;G. J. Duncan;M. D. Reid;G. J. Rucklidge;D. Henderson;P. Young;V. J. Russell;H. F. AminovP. Louis (2006). "A novel class of CoA-transferase involved in short-chain fatty acid metabolism in butyrate-producing human colonic bacteria." Microbiology 152: 179-185.

Chen, J. S.S. F. Hiu (1986). "Acetone butanol isopropanol production by Clostridium-Beijerinckii(Synonym: Clostridium Butylicum)." Biotechnol. Lett. 8: 371-376.

Chen, W. S.S. L. Lee (1997). "Optimization of medium composition for the production of glucosyltransferase by Aspergillus niger with response surface methodology." Enzyme Microb. Technol., 21(6): 436-440.

Cheng, S. S.;S. M. ChangS. T. Chen (2002). "Effects of volatile fatty acids on a thermophilic anaerobic hydrogen fermentation process degrading peptone." Water Sci. Tech. 46: 209-214.

Cheng, S. S.;S. M. ChangS. T. Chen (2002). Effects of volatile fatty acids on a thermophilic anaerobic hydrogen fermentation process degrading peptone. 2nd World Water Congress: Wastewater Treatment and Sludge Managment. 46: 209-214.

Chittibabu, G.;K. NathD. Das (2006). "Feasibility studies on the fermentative hydrogen production by recombinant Escherichia coli." Proc. Biochem. 41: 682-688.

174

Chu, C. F.;Y. Ebie;Y. InamoriH. N. Kong (2009). "Effect of hydraulic retention time on the hydrogen yield and population of Clostridium in hydrogen fermentation of glucose." J. of Enviromental Sciences (Beijing, China) 4: 424-428.

Classen, P.;J. Van Lier;C. Lopez;E. Van Niel;L. Sijtsma;A. Stams;S. De VriesR. Weusthuis (2000). "Utilization of biomass for the supply of energy carriers." Appl. Microbiol. Biotechnol. 52: 741-755.

Collet, C.;N. Adler;J. P. SchwitzguebelP. Peringer (2004). "Hydrogen production by Clostridium thermolacticum during continous fermentation of lactose." Int. J. Hydrog. Energy 29(14): 1479-1485.

Connor, M. R.J. C. Liao (2008). "Engineering of an Escherichia coli strain for the production of 3-methyl-1-butanol." Appl. Environ. Microbiol. 74(18): 5769-5775.

Copeland, A.;S. Lucas;A. Lapidus;K. Barry;J. C. Detter;T. Glavina;N. Hammon;S. Israni;E. Dalin;H. Tice;S. Pitluck;D. Sims;T. Brettin;D. Bruce;R. Tapia;J. Brainard;J. Schmutz;F. Larimer;M. Land;L. Hauser;N. Kyrpides;N. Mikhailova;G. Bennet;I. Cann;J. S. Chen;A. L. Contreras;D. Jones;E. Kashket;W. Mitchell;S. Stoddard;W. Schwarz;N. Qureshi;M. Young;Z. Shi;T. Ezeji;B. White;H. BlaschekP. Richardson (2009). "Clostridium beijerinckii NCIMB 8052, complete genome." http://www.ncbi.nlm.nih.gov/nuccore/CP000721.

Cornillot, E.;R. V. Nair;E. T. PapoutsakisP. Soucaille (1997). "The genes for butanol and acetone formation in Clostrium acetobutylicum ATCC 824 reside on a large plasmid whose loss leads to degeneration of the strain." J. Bacteriol 179: 5442-5447.

Cornillot, E.P. Soucaille (1996). "Solvent forming genes in Clostridia." Nature 380: 489.

Cotter, P. D.C. Hill (2003). "Surviving the acid test: response of gram-positive bacteria to low pH." Microb. Molec. Bio. Rev. 67(3): 429-453.

Crabbendam, P. M.;O. M. NeijsselD. W. Tempest (1985). "Metabolic and energetic aspects of the growth of Clostridium butyricum on glucose in chemostat culture." Archives of Microb. 142(4): 375-382.

Dabrock, B.;H. BahlG. Gottschalk (1992.). "Parameters affecting solvent production by Clostridium pasteurianum. ." Appl. Environ. Microbiol. 164: 35.

Danyluk, B.J. Kijowski (2001). "Influence of the lysozyme monomer on the growth of Clostridium tyrobutyricum bacteria." Przemysl Spozywczy 55(12): 16-19.

Das, D.T. N. Veziroglu (2001). "Hydrogen production by biological processes: a survey of literature." Int J Energy Res 26(13-28).

175

Desai, R. P.;L. M. Harris;N. E. WelkerE. T. Papoutsakis (1999). "Metabolic flux analysis elucidates the importance of the acid-formation pathways in regulating solvent production by Clostridium acetobutylicum." Metabo. Eng. 1: 206-213.

Desai, R. P.E. T. Papoutsakis (1999). "Antisense RNA strategies for metabolic engineering of Clostridium acetobutylicum " Appl. Environ. Microbiol. 65: 936-945.

Dharmadi, Y.;A. MurarkaR. Gonzalez (2006). "Anaerobic fermentation of glycerol by Escherichia coli: a new platform for metabolic engineering." Biotech. Bioeng. 94(5): 821-829.

Dos Santos, M. M.;V. Raghevendran;P. Kotter;L. OlssonJ. Nielsen (2004). "Manipulation of malic enzyme in Saccharomyces cerevisiae for increasing NADPH production capacity aerobically in different cellular compartments." Metabo. Eng. 6: 353-363.

Durot, M.;P. Y. BourguignonV. Schachter (2009). "Genome-scale models of bacterial metabolism: Reconstruction and applications." FEMS Microbio. Rev. 33(1): 164-190.

Dziedzak, J. D. (1986). "Biotechnology and flavor development: An industrial research perspective " Food Technology 40: 8-20.

Elam, C.;C. Gregoire-Padro;G. Sandrock;A. Luzzi;P. LindbladE. Hagen (2003). "Realizing the hydrogen future: the International Energy Agency's efforts to advance hydrogen energy technologies." Int. J. Hydrogen Energy 28: 601-607.

Ellman, G. L. (1959). Arch. Biochem. Biophysics 82(1): 70-77.

Energy Information Administration, U. S. D. o. E. (2008). "Weekly U.S. Retail Gasoline Prices." from see http://www.eia.doe.gov/oil_gas/petroleum/data_publications/ wrgp/mogas_home_page.html.

Fabiano, B.P. Perego (2002). "Thermeodynamic study and optimization of hydrogen production by Enterobacter aerogenes." Int. J. Hydrog. Energy 27: 149-156.

Fargione, J.;J. Hill;D. Tilman;S. PolaskyP. Hawthorne (2008). "Land cleaning and the biofuel carbon debt." Science 329(5867): 1235-1238.

Fleischmann, R. D.;M. D. Adams;O. White;R. A. Clayton;E. F. Kirkness;A. R. Kerlavage;C. J. Bult;J. F. Tomb;B. A. Dougherty;J. M. Merrick;K. Mckenney;G. Sutton;W. FitzHugh;C. Fields;J. D. Gocyne;J. Scott;R. Shirley;L. I. Liu;A. Glodek;J. M. Kelley;J. F. Weidman;C. A. Phillips;T. Spriggs;E. Hedblom;M. D. Cotton;T. R. Utterback;M. C. Hanna;D. T. Nguyen;D. M. Saudek;R. C. Brandon;L. D. Fine;J. L.

176

Fritchman;J. L. Fuhrmann;N. S. M. Geoghagen;C. L. Gnehm;L. A. McDonald;K. V. Small;C. M. Fraser;H. O. SmithJ. C. Venter (1995). "Wholegenome random sequencing and assembly of haemophilus." Science 5223(269): 507-512.

Fontaine, L.;I. Meynial-Salles;L. Girbal;X. H. Yang;C. CrouxP. Soucaille (2002). "Molecular characterization and transcriptional analysis of adhE2, the gene encoding the NADH-dependent aldehyde/alcohol dehydrogenase responsible for butanol production in alcohologenic cultures of Clostridium acetobutylicum ATCC 824." J. Bacteriol 184(3): 821-830.

Formanek, J.;R. MackieH. P. Blaschek (1997). "Enhanced butanol production by Clostridium beijerinckii BA101 grown in semidefined P2 medium containing 6% maltodextrin or glucose." Appl. Environ. Microbiol. 63: 2306-2310.

Frey, M. (2002). "Hydrogenases: Hydrogen-activationg enzymes." Chem. Bio. Chem. 3(2-3): 153-160.

Fritsch, M.;W. HartmeierJ. S. Chang (2008). "Enhancing hydrogen production of Clostridium butyricum using a cloumn reactor with square-structured ceramic fittings." Int. J. Hydrog. Energy 33(22): 6549-6557.

Fuhrmann, G.B. Volker (1992). "Regulation of glucose transport in Saccharomyces cerevisiae." J. Bacteriol 27: 1-15.

Gaertner, A. L.;H. Antelmann;R. Sapolsky;E. Ferrari;G. Chotani;B. MillerM. Hecher (2004). Qantitative proteomic and transcriptomic profiling during fermentation of pleiotropic Bacillus subtilis mutants. 227th ACS National Meeting. Anaheim, CA.

Goldberg, S.;J. Johnson;D. Busam;T. Feldblyum;S. Ferriera;R. Friedman;A. Halpern;H. Khouri;S. Kravitz;F. Lauro;K. Li;Y. Rogers;R. Strausberg;G. Sutton;L. Tallon;T. Thomas;E. Venter;M. FrazierJ. Venter (2006). "A sanger/pyrosequencing hybrid approach for the generation of high-quality draft asseblies of marine microbial genomes." P.N.A.S. 103(30): 11240-11245.

Gorwa, M. F.;C. CrouxP. Soucaille (1996). "Molecular characterization and transcriptional analysis of the putative hydrogenase gene of Clostridium acetobutylicum ATCC 824." J. Bacteriol 178(9): 2668-2675.

Gray, C. T.H. Gest (1975). "Biological formation of molecular hydrogen." Science 148: 186-192.

Green, E. M.G. N. Bennet (1997). "Genetic manipulation of acid and solvent formation in Clostridium acetobutylicum ATCC824." Biotech. Bioeng. 58: 215-221.

177

Green, E. M.;Z. L. Boynton;L. M. Harris;F. B. Rudolph;E. T. PapoutsakisG. N. Bennett (1996). "Genetic manipulation of acid formation pathways by gene inactivation in Clostridium acetobutylicum ATCC824." Microbiol 142: 2079-2086.

Gu, Z.;A. B. GlatzE. C. Glatz (1998). "Effects of propionic acid on propionibacteria fermentation." Enzyme Microb. Technol. 22: 13-18.

Guerout-Fleury, A. M.;K. Shazand;N. FrandsenP. Stragier (1995). "Antibiotic-resistance cassettes for Bacillus subtilis." Gene 167: 335-336.

Guettler, M. V.;M. K. JainB. K. Soni (1996). Process for making succinic acid, microorganisms for use in the process and methods of obtaining the microorganisms. US Patent 5504004. M. B. Institute.

Hallenbeck, P. C. (2005). "Fundamentals of the fermentative production of hydrogen." Water Science & Technology 52: 21-29.

Hallenbeck, P. C.J. R. Benemann (2002). "Biological hydrogen production: fundamentals and limiting processes." Int. J. Hydrogen Energy 27: 1185-1193.

Hanai, T.;S. AtsumiJ. C. Liao (2007). "Engineered synthetic pathway for isopropanol production in Escherichia coli." Appl. Environ. Microbiol. 73: 7814-7818.

Harris, L. M.;R. P. Desai;N. E. WelkerE. T. Papoutsakis (2000). "Characterization of recombinant strains of the Clostridium acetobutylicum butyrate kinase inactivation mutant: need for new phenomenological models for solventogenesis and butanol inhibition?" Biotech. Bioeng. 67: 1-11.

Hartimanis, M. G.S. Gatenbeck (1984). "Intermediary metabolism in clostridium acetobutylicum: levels of enzymes envolved in the fermentation of acetate and butyrate." Appl. Environ 47: 1277-1283.

Haug, M. (2004). Biofuels for transport; an international perspective. Paris, France, International Energy Agency.

Hawkes, F. R.;R. Dindale;D. L. HawkesI. Hussy (2002). "Sustainable fermentative biohydrogen: challenges for process optimization." Int J Energy Res 27: 1339-1347.

Hilgc-Rotmann, B.H. J. Rchm (1991). "Relationship between fermentation capability and fatty acid composition of free and immobilized Saccharomyces cerevisiae." Appl. Microbio. Biotech. 34: 502-508.

Hsu, T. A.;M. R. LadischG. T. Tsao (1980). "Alcohol from cellulose." Chem. Technol. 10(5): 315-319.

178

Huang, W. C.;D. E. RameyS. T. Yang (2004). "Continuous production of butanol by Clostridium acetobutylicum immobilized in a fibrous bed bioractor " Appl. Biochem. Biotechnol. 115: 887-898.

Huang, Y. L., Wu, Z.T., Zhang, L.K., Cheung, C.M., Yang, S.T. (2002). "Production of carboxylic acids from hydrolyzed corn meal by immobilized cell fermentation in fibrous-bed bioreactor." Bioresour Technol 82(1): 51-59.

Huang, Y. L.;K. Mann;J. M. NovakS. T. Yang (1998). "Acetic acid production from fructose by Clostridium formicoaceticum immobilized in a fibrous-bed bioreactor." Biotech. Prog. 14: 800-806.

Huang, Y. L.S. T. Yang (1998). "Acetate production from whey lactose using co-inmmobilized cells of homolactic and homoacetic bacteria in a fibrous-bed bioreactor." Biotech. Bioeng. 60: 499-507.

Huber, G. W.;S. IborraA. Corma (2006). "Synthesis of transportation fuels from biomass: chemistry, catalysis, and engineering." Chem. Rev. 106: 4044-4098.

Ikeda, M.S. Nakagawa (2003). "The corynebacterium glutamicum genome: features and impacts on biotechnological processes." Appl. Microbiol. Biotechnol. 62: 99-109.

Ingram, L. O. (1976). "Adaptation of membrane lipids to alcohols." J. Bacterial 125: 670-678.

Ito, T.;Y. Nakashimada;K. Senba;T. MatsuiN. Nishio (2005). "Hydrogen and ethanol production from glecerol-containing wastes discharged after biodiesel manufacturing process." J. Biosci. Bioeng. 100(3): 260-265.

Ivey, D. M.L. G. Ljungdahl (1986). "Purification and characteristics of the F1-ATPase from Clostridium thermoaceticum." J. Bacterial 165: 252-257.

Jiang, L.;J. F. Wang;S. Z. Liang;X. N. Wang;P. L. CenZ. N. Xu (2009). "Butyric acid fermentation in a fibrous bed bioreactor with immoblilized Clostridium tyrobutyricum from cane molasses." Bioresour Technol 100(13): 3403-3409.

Jiraskova, A.;L. Vitek;J. Fevery;T. RumlP. Branny (2005). "Rapid protocol for electroporation of Clostridium perfeingens." J. Microbiol Methods 62: 125-127.

Jo, J. H.;D. S. Lee;D. H. ParkJ. M. Park (2008). "Statistical optimization of key process variables for enhanced hydrogen production by newly isolated Clostridium tyrobutyricum JM1." Int J Energy Res 33(19): 5176-5183.

179

Jo, J. H.;D. S. LeeJ. M. Park (2008). "The effects of pH on carbon material and energy balances in hydrogen-producing Clostridium tyrobutyricum JM1." Bioresour Technol 99(17): 8485-8491.

Johnson, J.B. Francis (1975). "Taxonomy of the clostridia: ribosomal ribonucleic acid homologies among species." J. Gen. Microbial. 88: 229-244.

Johnson, J. L.J. S. Chen (1995). "Taxonomic relationships among strains of Clostridium acetobutylicum and other phenotypically similar organisms." FEMS Microbio. Rev. 17: 233-240.

Jones, D. T.S. Keis (1995). "Origins and relationships of industrial solvent-producing clostridial stains." FEMS Microbio. Rev. 17: 223-232.

Jones, D. T.D. R. Woods (1986). "Acetone-butanol fermentation revisited." Microb. Rev. 50(4): 484-524.

Jones, S. F.;G. M. EvansK. P. Galvin (1999). "The cycle of bubble production from a gas cavity in a supersaturated solution." Adv. Colloid and Inerface Sc. 80(1): 51-84.

Jungermann, K.;G. Leimenstoll;E. RupprechrtR. K. Thauer (1971). "Demonstration of NADH-ferrodoxin reductase in two saccharolytic clostridia." Archives fuer Mikrobiologie 80(4): 370-372.

Kakuda, H.;K. Hosono;K. ShiroshiS. Ichihara (1994). "Identification and characterization of the ackA (acetate kinase A)-pta (phosphotransacetylase) operon and complementation analysis of acetate utilizaiton by an ackA-pta deletion mutant of Escherichia coli." J. Biochem. 116: 916-922.

Kalscheuer, R.;T. StoeltingA. Steinbuechel (2006). "Microdiesel: Escherichia coli engineered for fuel production." Microbiology (Reading, United Kingdom) 152(9): 2529-2536.

Karube, I.;N. Urano;T. Yamada;H. HirochikaK. Sakaguchi (1983). "Cloning and expression of the hydrogenase gene from Clostridium butyricum in Escherichia coli." FEBS Lett. 158: 119-122.

Kataoka, N.;A. MiyaK. Kiriyama (1997). "Studies on hydrogen production by continuous culture system of hydrogen producing anaerobic bacteria." Water Science & Technology 36: 41-47.

Kawagoshi, Y.;N. Hino;A. Fujimoto;M. Nakao;Y. Fujita;S. SugimuraK. Furukawa (2005). "Effect of inoculum conditioning on hydrogen fermentation and pH effect on bacterial community relevant to hydrogen production." J. Biosci. Bioeng. 100(5):

180

524-530.

Kayser, A.;J. Weber;V. HechtU. Rinas (2005). "Metabolic flux analysis of Escherichia coli in glucose-limited continuous culture. I. Growth-rate dependent metabolic efficiency at steady state." Microbiology 151: 693-706.

Keis, S.;C. F. Bennett;V. K. WardD. T. Jones (1995). "Taxonomy and phylogeny of industrial solvent-producing clostridia." int. J. Syst. Bacterial. 45: 693-705.

Keis, S.;R. ShaheenD. T. Jones (2001). "Emended descriptions of Clostridium acetobutylicum and Clostridium beijerinckii and descriptions of Clostridium saccharoperbutylacetonicum sp. nov. and Clsotridium saccharobutylicum sp. nov." Int. J. Syst. Evo. Microb. 51: 2095-2103.

Keshav, A.;K. L. WasewarS. Chand (2009). "Extraction of acrylic, propionic, and butyric acid using Aliquat 336 in oleyl alcohol: equilibria and effect of temperature." Industrial & Engineering Chemistry Research 48(2): 888-893.

Khanal, S. K.;W. H. Chen;L. LiS. W. Sung (2004). "Biological hydrogen production: effects of pH and intermediate products." Int. J. Hydrog. Energy 29 1123-1134.

Killeffer, D. H. (1927). "Butanol and acetone from corn." J. Indus. Eng. Chem. 19: 46-50.

Klass, D. L. (1998). Biomass for renewable energy, fuels and chemicals. San Diego, CA, Academic Press.

Klijn, N.;F. Nieuwenhof;J. Hoolwerf;V. d. W. C.A. Weerkamp (1995). "Identification of Clostridium tyrobutyricum as the causative agent of late blowing in cheese by species-specific PCR amplification." Appl. and Enviro. Microbiol. 61(8): 2919-2924.

Kraemer, J. T.D. M. Bagley (2007). "Improving the yield from fermentative hydrogen production." Biotechnol. Lett. 29: 685-695.

Kroschwitz, J. I.M. E. Howe-Grant (1997). Sugar to thin films. Kirk-Othermer Encyclopedia of Chemical Technology. Fourth Edition. 23: 1118.

Kumar, N.D. Das (2000). "Enhancement of hydrogen production by Enterobacter cloacae IIT-BT08." Proc. Biochem. 35: 589-593.

Kumar, N.D. Das (2001). "Continuous hydrogen production by immobilized enterobacter cloacae IIT-BT 08 using lignocellulosic materials as solid matrices." Enzyme Microbiol. Technology 29: 280-287.

Kumar, N.;A. GhoshD. Das (2001). "Redirection of biochemical pathways for the

181

enhancement of H2 production by Enterobacter cloacae." Biotechnol. Lett. 23: 537-541.

Kumar, P.;D. M. Barrett;M. J. DelwicheP. Strove (2009). "Methods for pretreatment of ligncellulosic biomass for effeient hydrolysis and biofuel production." Industrial & Engineering Chemistry Research 48(8): 3713-3729.

Ladygina, N.;E. G. DedyukhinaM. B. Vainshtein (2006). "A review on micorbial systhesis of hydrocarbons." Proc. Biochem. 41(5): 1001-1014.

Lagunas, R. (1993). "Sugar transport in Saccharomyces cerevisiae " FEMS Microbio. Rev. 104: 229-242.

Larger, S. T.;S. Long;J. D. Santangelo;D. T. JonesD. R. Woods (1985). "Immobilized Clostridium acetobutylicum P262 mutants for solvent production." Appl. Environ. Microbiol. 50: 477-481.

Latimer, M. T.J. G. Fery (1993). "Cloning, sequence analysis, and hyperexpression of the genes encoding phosphotransacetylase and acetate kinase from Methanosarcina thermophila." J. Bacterial 175: 6822-6829.

Lay, J. J. (2000). "Modeling and optimization of anaerobic digested sludge converting starch to hydrogen. ." Biotechnol. Bioeng. 68: 269-278.

Lee, J.;W. J. Mitchell;M. TangneyH. P. Blaschek (2005). "Evidence for the presence of an alternative glucose transport system in Clostridium beijerinckii NCIMB 8052 and the solvent-hyperproducing mutant BA101." Appl. Environ. Microbiol. 71(6): 3384-3387.

Lee, P.;M. Bulls;J. HolmesJ. W. Barrier (1997). "Hybrid process for the conversion of lignocellulosic materials." Appl. Biochem. Biotechnol. 66: 1-23.

Lee, S. M.;M. O. Cho;C. H. Park;Y. C. Chung;J. H. Kim;B. I. SangY. Um (2008). "Continuous butanol production using suspended and immobilized Clostridium beijerinckii NCIMB 8052 with supplementary butyrate." Energy & Fuels 22(3459-3464).

Lee, Y. J. (2004). "Effects of pH on continuous hydrogen fermentation." Hangug Hwangyeong bogeon Haghoeji 30(2): 149-153.

Lee, Y. J.;T. MiyaharaT. Noike (2001). "Effect of iron concentration on hydrogen fermentation." Bioresour Technol 80(3): 227-231.

Lee, Y. J.;T. MiyaharaT. Noike (2002). "Effect of pH on microbial hydrogen fermentation." J. Chem. Technol. Biotechnol. 77(6): 694-698.

Lemke, H. (1963). "New catalyst cycle in the oxo synthesis." Genie Chimique 89:

182

118-120.

Lenaz, G.;C. G. ParentiA. M. Sechi (1975). "Lipid-protein interactions in mitochondria: changes in mitochondrial adenosine triphosphatase activity induced by n-butyl alcohol." Arch. Biochem. Biophys 167: 72-79.

Leonard, E.;Z. L. FowlerM. Koffas (2007). "Metabolic engieering." Cell Engineering 5: 301-359.

Levin, D. B.;L. PittM. Love (2004). "Biohydrogen production:prospects and limitations to practical application." Int J Energy Res 29: 173-185.

Li, M. Z.S. J. Elledge (2007). "Harnessing homologous recombination in vitro to generate recombinant DNA via SLIC." Nature Methods 4(3): 251-256.

Liao, J. C.W. Higashide (2008). "Metabolic engineering of next-generation biofuels." Chem. Eng. Prog. 104(8): S19-S23.

Lienhardt, J.;J. Schripsema;N. QureshiH. P. Laschek (2002). "Butanol production by Clostridium beijerinckii BA101 in an immobilized cell biofilm reactor." Appl. Biochem. Biotechnol. 98: 591-598.

Lin, S. K.;C. Y. Du;A. Koutinas;R. H. WangC. Webb (2008). "Substrate and product inhibition kinetics in succinic acid production by Actinobacillus succinogenes." Biochem. Eng. J. 41: 128-135.

Linden, J. C.A. Moreira (1982). "Anaerobic production of chemicals." Biological basis for new developments in biotechnology 25: 377-403.

Lipmann, F.L. C. Tuttle (1945). "A specific micromethod for the determination of acyl phosphate." J. Bio. Chem. 159: 21-28.

Liu, X.;Y. ZhuS. T. Yang (2006). "Construction and characterization of ack deleted mutant of Clostridium tyrobutyricum for enhanced butyric acid and hydrogen production." Biotech. Prog. 22: 1265-1275.

Liu, X. G.S. T. Yang (2006). "Kinetics of butyric acid fermentation of glucose and xylose by Clostridium tyrobutyricum wild type and mutant." Proc. Biochem. 41(4): 801-808.

Liu, X. G.;Y. ZhuS. T. Yang (2006). "Butyric acid and hydrogen production by Clostridium tyrobutyricum ATCC25755 and mutants." Enzyme Microb. Technol. 38: 521-528.

Liu, Y. P.;P. Zheng;Z. H. Sun;Y. Ni;J. J. DongL. L. Zhu (2008). "Economical succinic

183

acid production from cane molasses by Actinobacillus succinogenes." Bioresour Technol 99: 1736-1742.

Maeda, T.;V. Sanchez-TorresT. K. Wood (2007). "Enhanced hydrogen production from glucose by a metabolically-engineered Escherichia coli." Appl. Environ. Microbiol. 77(4): 879-890.

Maeda, T.;V. Sanchez-TorresT. K. Wood (2008). "Metabolic engineering to enhance bacterial hydrogen production." Microb. Biotech. 1(1): 30-39.

Magnusson, L.;R. Islam;R. Sparling;D. LevinN. Cicek (2008). "Direct hydrogen production from cellulosic waste materials with a single-step dark fermentation process." Int. J. Hydrog. Energy 33(20): 5398-5403.

Mahyudin, A. R.;Y. Furutani;Y. Nakashimada;T. KakizonoNishio (1997). "Enhanced hydrogen production n altered mixed acid fermentation of glucose by Enterobacter aerogenes." J. Ferm. Bioeng. 83(4): 358-363.

Maiorov, D. M.;D. V. MushenkoM. P. Vysotskii (1971). "Production of alcohols, acids, and aldehydes based on petroleum." Chem. Technol. of Fuels 7: 495-502.

Margulies, M.;M. Egholm;W. E. Altman;S. Attiya;J. S. Bader;L. A. Bemben;J. Berka;M. S. Braverman;Y. J. Chen;Z. Chen;S. B. Dewell;L. Du;J. M. Fierro;X. V. Gomes;B. C. Godwin;W. He;S. Helgesen;C. H. Ho;G. P. Irzyk;S. C. Jando;M. L. Alenquer;T. P. Jarvie;K. B. Jirage;J. B. Kim;J. R. Knight;J. R. Lanza;J. H. Leamon;S. M. Lefkowitz;M. Lei;J. Li;K. L. Lohman;H. Lu;V. B. Makhijani;K. E. McDade;M. P. McKenna;E. W. Myers;E. Nickerson;J. R. Nobile;R. Plant;B. P. Puc;M. T. Ronan;G. T. Roth;G. J. Sarkis;J. F. Simons;J. W. Simpson;M. Srinivasan;K. R. Tartaro;A. Tomasz;K. A. Vogt;G. A. Volkmer;S. H. Wang;Y. Wang;M. P. Weiner;P. Yu;R. F. BegleyJ. M. Rothberg (2005). "Genome sequencing in microfabricated high-density picolitre reactors." Nature 437(7057): 376-380.

Mari, F.;G. Oise;C. CrouxP. Soucille (1996). "Molecular characterization and transcriptional analysis of the putative hydrogenase gene of Clostridium acetobutylicum ATCC824." J. Bacterial 178(9): 2668-2675.

Matsuyama, H.;H. YamamotoE. Nakano (1989). "Cloning, expression, and nucleotide sequences of the Escherichia coli. K-12 ackA gene." J. Bacterial 171: 77-80.

Mermelstein, L. D.E. T. Papoutsakis (1993). "In vivo methylation in Escherichia coli by the Bacillus subtilis phage_3TI methyltransferase to protect plasmids from restriction upon transformation of Clostridium acetobutylicum ATCC 824." Appl. Environ. Microbiol. 59: 1077-1081.

184

Meyer, J.J. Gagnod (1991). "Primary structure of hydrogenase I from Clostridium pasteurianumt." Biochem. 30: 9697-9704.

Michalodimitrakis, K.M. Isalan (2009). "Engineering prokaryotic gene circuits." FEMS Microbio. Rev. 33(1): 27-37.

Michel-Savin, D.;R. MarchalJ. P. Vandecasteele (1990). "Butyrate production in continuous culture of Clostridium tyrobutyricum: effect of end-product inhibition." Appl. Microbio. Biotech. 33(2): 127-131.

Michel-Savin, D.;R. MarchalJ. P. Vandecasteele (1990). "Control of the selectivity of butyric acid production and improvement of fermentation performance with Clostridium tyrobutyricum." Appl. Microbio. Biotech. 32(4): 387-392.

Michel-Savin, D.;R. MarchalP. Vandecasteele (1990). "Butyric fermentation; Metabolic behavior and production performance of Clostridium tyrobutyricum in a continuous culture with cell recycle." Appl. Microbiol. Biotechnol. 34(172-177).

Minteer, S. D. (2006). Alcoholic fuels: An overview. Boca Raton, CRC Press.

Mishra, J.;S. Khurana;N. Kumar;A. K. GhoshD. Das (2004). "Molecular cloning, characterization, and overexpression of a novel [Fe]-hydrogenase isolated from a high rate of hydrogen producing Enterobacter cloacae IIT-BT08." BioChem. Biophys. Res. Commun. 324: 679-685.

Mitchell, W. J. (1998). "Physiology of carbohydrate to solvent conversion by clostridia." Adv. Microb. Physiol. 39: 31-130.

Moreira, A. R.;D. C. UlmerJ. C. Linden (1981). "Butanol toxicity in the butyric fermentation." Biotech. Bioeng. Syrup 11: 567-579.

Mosier, N. S.;C. Wyman;B. Dale;R. Elander;Y. Y. Lee;M. HoltzappleM. R. Ladisch (2005). "Features of promising technologies for pretreatment of lignocellulosic biomass." Bioresour Technol 96: 673-686.

Mullis, K.;F. Faloona;S. Scharf;R. Saiki;G. HornH. Erlich (1986). "Specific enzymatic amplification of DNA in vitro: the polymerase chain reaction." Cold Spring Harb. Symp. Quant Biol. 51(Pt.1): 263-273.

Nair, R. V.;G. N. BennettE. T. Papoutsakis (1994). "Molecular characterization of an aldehyde/alcohol dehydrogenase gene from C. acetobutylicum ATCC 824." Journal of Bacteriology 176: 871.

Nair, R. V.E. T. Papoutsakis (1994). "Expression of plasmid-encoded aad in clostridium

185

M5 restores vigorous butanol production." Journal of Bacteriology 176: 5843.

Nandi, R.;S. DeyS. Sengupta (2001). "Thiosulphate improves yield of hydrogen production from glucose by the immobilized formate hydrogenlyase system of Escherichia coli." Biotech. Bioeng. 75: 492-494.

Nath, K.D. Das (2004). "Improvement of fermentative hydrogen production: various approaches." Appl. Microbio. Biotech. 65: 520-529.

News, C. E. Production report. Chem. Eng. News, ACS. July 11, 2005 83: 67-76.

Noike, T.;H. Takabatake;O. MizunoM. Ohba (2002). "Inhibition of hydrogen fermentation of organic wastes by lactic acid bacteria." Int J Energy Res 27: 1367-1371.

Nolling, J.;G. BretonM. V. Omelchenko (2001). "Genome sequence and comparative analysis of the solvent-producing bacterium Clostridium acetobutylicum." J. Bacteriol 183: 4823-4838.

Norton, S.;C. LacroixJ. C. Vuillemard (1994). "Kinetic study of continuous whey permeate fermentation by immobilized Lactobacillus helveticus for lactic acid production." Enzyme Microb. Technol. 16(6): 457-466.

O'Sullivan, E.S. Condon (1999). "Relationship between acid tolerance, cytoplasmic pH, and ATP and H+-ATPase levels in chemist cultures of Lactococcus lactis." Appl. Environ. Microbiol. 65: 2287-2293.

Ozkan, P.;B. Sariyar;F. O. Utkur;U. AkmanA. Hortacsu (2005). "Metabolic flux analysis of recombinant protein overproduction in Escherichia coli." Biochem. Eng. J. 22: 167-195.

Parro, V.M. Moreno-Paz (2003). "Gene function analysis in environmental isolates: the nif regulon of the strict iron oxidazing bacterium Leptospirillum ferrooxidans." Proc. Natl. Acad. Sci. USA 100: 7883-7888.

Patel, G. B.B. J. Agnew (1988). "Growth and butyric acid production by Clostridium populeti." Achives of Microb. 150(3): 267-271.

Penfold, D. W.;C. F. ForsterL. E. Macaskie (2003). "Increased hydrogen production by Escherichia coli strain HD701 in comparison with the wild-type parent strain MC4100." Enzyme Microb. Technol. 33: 185-189.

Penfold, D. W.;F. SargentL. E. Macaskie (2006). "Inactivation of the Escherichia coli K-12 twin-arginine translocation system promotes increased hydrogen production." FEMS Microbio. Lett. 262: 135-137.

186

Pennisi, E. (2005). "Biochemistry, cut-rate genomes on the horizon?" Science 309(5736): 862.

Perez, M.;K. Luyten;R. Michel;C. RiouB. Blondin (2005). "Analysis of Saccharomyces cerevisiae hexose carrier expression during wine fermentation: both low- and high-affinity Hxt transporters are expressed." FEMS Yeast Res. 5: 351-361.

Perfil'ev, G. D.;A. V. GudkovN. A. II'inskaya (1978). "Effect of iron ions on the growth and metabolism of Clostridium tyrobutyricum which ferments butyric acid in cheese." Trudy, Vasesoyuznyi Nauchno-Issledovatel'skii Institut Maslodel'noi i Syrodel'noi Promyshlennosti 23: 72-77.

Perlack, R. D.;L. L. Wright;A. Turhollow;R. L. Graham;B. StokesD. C. Erbach (2005). Biomass as feedstock for a bioenergy and bioproducts industry: The technical feasibility of a billion-ton annual supply. Oak Ridge, TN, Oak Ridge National Laboratory. DOE/GO-102995-2135.

Petitdemange, H.;J. M. Bengone;J. L. BergereR. Gay (1973). "Regulation of the NAD+ and NADP+-ferredoxin reductase activities in a Clostridium of the butyric acid group: Clostridium tyrobutyricum." Lab. Chim. Biol. , Univ. Nancy-I, Nancy Fr. Biochimie 55(10): 1307-1310.

Pfleger, B. F.;D. J. Pitera;C. D. SmolkeJ. D. Keasling (2006). "Combinatorial engineering of intergenic regions in operons tunes expression of multiple genes." Nature Biotech. 24(8): 1027-1032.

Phillips-Jones, M. K. (1995). "Introduction of recominant DNA into Clostridium spp. ." Methods Mol. Biol. 47: 227-235.

Playne, M. J. (1985). "Determination of ethanol, volatile fatty acids, lactic and succinic acids in fermentation liquids by gas chromatography." J. Sci. of Food Agri. 36(8): 638-644.

Pouillart, P. R. (1998). "Role of butyric acid and its derivatives in the treatment of colorectal cancer and homoglobinopathies." Life Sci. 63(20): 1739-1760.

Price, N. D.;J. L. ReedB. O. Palsson (2004). "Genome-scale models of microbial cells: evaluating the consequences of constraints." Nature Rev. Microb. 2(11): 886-897.

Pryde, E. H. (1978). "Carboxylic acids-survey." Kirk-Othermer Encycl. Chem. Technol. 3rd Ed. 4: 814-834.

Rachman, M. A.;Y. Furutani;Y. Nakashimada;T. KakizonoN. Nishio (1997). "Enhanced

187

hydrogen production in mixed acid fermentation of glucose by Enterobacter aerogenes." J. Fermen. and Bioeng. 83: 358-363.

Ramey, D. E. (1998). "Continuous two stages, dual path anaerobic fermentation of butanol and other organic solvents using two different strains of bacteria." US patent 5753474.

Ren, N. Q.;J. Ding;L. Ding;M. Liu;Y. F. LiH. X. Bao (2004). "Effect of Cu2+ concentration on hydrogen fermentation by mixed culture." J. Harbin Institute of Tech.(Eng. Ver.) 11(1): 11-16.

Rephaeli, A.;R. ZhukA. Nudelman (2000). "Prodrugs of butyric acid from bench to bedside: synthetic design, mechanisms of action, and clinical applications." Drug Development Resear. 50(3/4): 379-391.

Riebeling, V.K. Jungermann (1976). "Properties and function of Clostridial membrane ATPase." Biochem. Biophys. Acta 430(3): 434-444.

Rose, I. A. (1955). "Acetate kinase of bacteria (acetokinase)." Methods Enzymol. 1: 591-595.

Saint-Amans, S.;L. Girbal;J. Andrade;K. AhrensP. Soucaolle (2001). "Regulation of carbon and electron flow in Clostridium butyricum VPI3266 grown in glucose-glycerol mistures." J. Bacteriology 183: 1748-1754.

Sanchez, A. M.;J. Andrews;I. Hussein;G. N. BennettK. Y. San (2006). "Effect of overexpression of a soluble pyridine nucleotide transhydrogenase (UdhA) on the production of poly(3-hydroxybutyrate) in Escherichia coli." Biotechnol. Prog. 22: 420-425.

Sanger, F.;G. M. Air;B. G. Barrell;N. L. Brown;A. R. Coulson;C. A. Fiddes;C. A. Hutchison;P. M. SlocombeM. Smith (1977). "Nucleotide sequence of bacteriophage phi X174 DNA." Nature 265(5596): 687-695.

Sanger, F.;J. E. Donelson;A. R. Coulson;H. KosselD. Fischer (1973). "Use of DNA polymerase I primed by a synthetic oligonucleotide to determine a nucleotide sequence in phage f1 DNA." Proc. Natl. Acad. Sci. U.S.A. 70(4): 1209-1213.

Schoondermark-Stolk, S. A.;M. Jansen;J. H. Veurink;A. J. Verkleij;C. T. Verrips;G. J. Euverink;J. BoonstraL. Dijkhuizen (2006). "Rapid identification of target genes for 3-methyl-1-butanol production in Saccharomyces cerevisiae." Appl. Microbio. Biotech. 70: 237-246.

Scott, P. T.J. I. Rood (1989). "Electroporation-mediated transformation of

188

lysostaphin-treated Clostridium perfringens." Gene 82: 327-333.

Sebaihia, M.;M. W. Peck;N. P. Minton;N. R. Thomson;M. T. Holden;W. J. Mitchell;A. T. Carter;S. D. Bentley;D. R. Mason;L. Crossman;C. J. Paul;A. Ivens;M. H. Wells-Bennik;I. J. Davis;A. M. Cerdeno-Tarraga;C. Churcher;M. A. Quail;T. F. Chillingworth;T. Feltwell;A. Fraser;I. Goodhead;Z. Hance;K. Jagels;N. Larke;M. Maddison;S. Moule;K. Mungall;H. Norbertczak;E. Rabbinowitsch;M. Sanders;M. Simmonds;B. WhiteS. P. Whithead, J. (2007). "Genome sequence of a proteolytic (Group I) Clostridium botulinum strain Hall A and comparative analysis of the clostridial genomes." Genome Res. 17(1082-1092).

Sebaihia, M.;B. W. Wren;P. Mullany;N. F. Fairweather;N. Minton;R. Stabler;N. R. Thomson;A. P. Roberts;A. M. Cerdeno-Tarraga;H. W. Wang;M. T. Holden;A. Wright;C. Churcher;M. A. Quail;S. Baker;N. Bason;K. Brooks;T. Chillingworth;A. Cronin;P. Davis;L. Dowd;A. Fraser;T. Feltwell;Z. Hance;S. Holroyd;K. Jagels;S. Moule;K. Mungall;C. Price;E. Rabbinowitsch;S. Sharp;M. Simmonds;K. Stevens;L. Unwin;S. Whithead;B. Dupuy;G. Dougan;B. BarrellJ. Parkhill (2006). "The multidrug-resistant human pathogen Clostridium difficile has a highly mobile, mosaic genome." Nature Gene. 38: 779-786.

Sentheshanmuganathan, S. (1960). "Mechanism of the formation of higher alcohols from amino acids by Saccharomyces cerevisiae." Biochem. J. 74: 568-576.

Shapovalova, O. I.L. A. Ashkinazib (2008). "Biobutanol: Biofuel of second generation." Russian J. Appl. Chem. 81(12): 2232-2236.

Shendure, J.;G. J. Porreca;N. B. Reppas;X. Lin;J. P. McCutcheon;A. M. Rosenbaum;M. D. Wang;K. Zhang;R. D. MitraG. M. Church (2005). "Accurate multiplex polony sequencing of an evolved bacterial genome." Science 309(5741): 1728-1732.

Shetty, R. P.;D. EndyT. F. Knight (2008). "Engineering biobrick vectors from biobrick parts." J. Bio. Eng. 2(5): doi:10.1186/1754-1611-2-5.

Shimizu, T.;K. Ohtani;H. Hirakawa;K. Ohshima;A. Yamashita;T. Shiba;N. Ogasawara;M. Hattori;S. KuharaH. Hayashi (2002). "Complete genome sequence of Clostridium perfringens, an anaerobic flesh-eater." PNAS 99(2): 996-1001.

Shin, J. H.;J. H. Yoon;E. K. Ahn;M. S. Kim;S. J. SimT. H. Park (2007). "Fermentative hydrogen production by the newly isolated Enterobacter asburiae SNU-1." Int. J. Hydrog. Energy 32: 192-199.

Shine, J.L. Dalgarno (1974). "The 3'-terminal sequence of E. coli 16S ribosomal RNA: complientartiy to nonsense triplets and bibosome bonding site." Proc. Natl. Acad. Sci. USA 71: 1342.

189

Shinto, H.;Y. Tashiro;M. Yamashita;G. Kobayashi;T. Sekiguchi;T. Hanai;Y. Kuriya;M. OkamotoK. Sonomoto (2007). "Kinetic modeling and sensitivity analysis of acetone-butanol-ethanol production." J. Biotech. 131: 45-56.

Sillers, R.;M. Al-HinaiE. T. Papoutsakis (2009). "Aldehyde-alcohol dehydrogenase and/or thiolase overexpression coupled with CoA transferase downregulation lead to higher alcohol titers and selectivity in Clostridium acetobutylicum fermentations." Biotechnol. Bioeng. 102(1): 38-49.

Silva, E. M.S. T. Yang (1995). "Kinetics and stability of a fibrous-bed bioreactor for continuous production of lactic acid from unsupplemented acid whey." J. Biotechnol. 41: 59-70.

Soucaille, P.G. Goma (1986). "Acetonobutylic fermentation by Clostridium acetobutylicum ATCC 824: Autobacteriocin production, properties and effects." Current Microbiol. 13: 163-169.

Stephanopoulos, G. N.;A. A. AristidouJ. Nielsen (1998). Metabolic Engineering: Principles and Methodologies. San Diego, USA, Academic Press.

Suresh, B. (2003). Acetic acid, SRI International. CEH Report 602.5000.

Suwannakham, S.;Y. HuangS. T. Yang (2006). "Construction and characterization of ack knock-out mutants of Propionibacterium acidpropionici for enhanced propionic acid fermentation." Biotech. Bioeng. 94(2): 383-395.

Tangney, M.;A. Galinier;J. DeutscherW. J. Mitchell (2003). "Analysis of the elements of catabolite repression in Clostridium acetobutylicum ATCC 824." J. Microb. Biotech. 6: 6-11.

Tanisho, S.;Y. SuzukiN. Wakao (1987). "Fermentative hydrogen evolution by Entebacter aerogenes E. 82005." Int. J. Hydrogen Energy 12(9): 623-627.

Tejavadi, S.M. Cheryan (1988). "Downstream processing of lactic acid whey permeate fermentation broths by hollow fiber ultrafiltration." Appl. Biochem. Biotechnol. 19(1): 61-70.

Thauer, R. K.;K. JungermannK. Decker (1977). "Energy conservation in chemotrophic anaerobic bacteria." Bioteriological Reviews 41(1): 80-100.

Thurell, K. E.G. Sjostrom (1952). "Effect of nitrite and nitrate on butyric acid fermentation in cheese." Svenska Mejeritdningen 44: 511-514, 523-526.

190

Tishkov, V. I.V. O. Popov (2006). "Protein engineering of formate dehydrogenase." Biomol. Eng. 23: 89-110.

Tokgoz, S.;D. HayesT. H. Yu (2008). "Use of U.S. Croplands for biofuels increases greenhouse gases through emissions from land use change." Science 319(5867): 1238-1240.

Tolvanen, K.;P. Koskinen;H. Raussi;A. Ylikoski;I. Hemmilae;V. SantalaM. Karp (2008). "Profiling the hydA gene and hydA gene transcript levels of Clostridium butyricum during continuous, mixed-culture hydrogen fermentation." Int. J. Hydrog. Energy 33(20): 5416-5421.

Tomas, C. A.;K. V. Alsaker;H. P. Bonarius;W. T. Henderiksen;H. Yang;J. A. Beamish;C. J. ParedesE. T. Papoutsakis (2003). "DNA array-based transcriptional analysis of asporogenous, nonsolventogenic Clostridium acetobutylicum strains SKO1 and M5." J. Bacteriol 185: 4539-4547.

Trinh, C. T.;A. WlaschinF. Srienc (2009). "Elementary mode analysis: A useful metabolic pathway analysis tool for characterizing cellular metabolism." Appl. Microbio. Biotech. 81(5): 813-826.

Tummala, S. B.;S. G. JunneE. T. Papoutsakis (2003). "Antisense RNA downregulation of coenzyme A transferase combined with alcohol-aldehyde dehydrogenase overexpression leads to predominantly alcohologenic Clostridium acetobutylicum fermentations " J. Bacteriol 185: 3644-3653.

Turner, J.;G. Sverdrup;M. K. Mann;P. C. Maness;B. Kroposki;M. Ghirardi;R. J. EvansD. Blake (2008). "Renewable hydrogen production." Int J Energy Res 32(5): 379-407.

Urbance, S. E.;A. L. Pometto III;A. A. DiSpititoY. Denli (2004). "Evaluation of succinic acid continuous and repeat-batch biofilm fermentation by Actinobacillus succinogenes using plastic composite support bioreactors." Appl. Microbio. Biotech. 65: 664-670.

Valdez-Vazquez, I.;R. Sparling;D. Risbey;N. Rinderknecht-SeijasH. M. Poggi-Varaldo (2005). "Hydrogen generation via anaerobic fermentation of paper mill wastes." Bioresour Technol 96(17): 1907-1913.

Valentine, R. C.R. S. Wolfe (1960). "Purification and role of phosphotransbutyrylase." J. Bio. Chem. 235: 1948-1952.

Van Groenestijn, J. W.;J. H. Hazewinkel;M. NienoordP. J. Bussmann (2002). "Energy aspects of biological hydrogen production in high rate bioreactors operated in the thermophilic temperature range." Int. J. Hydrog. Energy 27(11): 1141-1147.

191

Vardar-Schara, G.;T. MaedaT. K. Wood (2008). "Metabolically engineered bacteria for producing hydrogen via fermentation." microbiol Biotech. 1(2): 107-125.

Varma, A.B. O. Palsson (1994). "Metabolic flux balancing- basic concepts scientific and practical use." Bio-Technol. 12: 994-998.

Vignais, P. M.;B. BilloudJ. Meyer (2001). "Classification and phylogeny of hydrogenases." FEMS Microbio. Rev. 25(4): 455-501.

Vignais, P. M.A. Colbeau (2004). "Molecular biology of microbial hydrogenases." Curr. Issues in Mol. Bio. 6(2): 159-188.

Vos, E. A. (1948). "The influence of potassium nitrate on butyric acid fermentation of cheese." Netherlands Milk and Dairy Journal 2(223-245).

Wackett, L. P. (2008). "Microbial-based motor fuels: science and technology." Microb. Biotech. 1(3): 211-225.

Walsh, M. C.;H. P. Smits;M. ScholteK. Van Dam (1994). "Affinity of glucose transport in Saccharomyces cerevisiae is modulated during growth on glucose." J. Bacteriol 176: 953-958.

Walter, K.;R. Nair;J. CaryG. Bennett (1993). "Sequence and arrangement of two genes of the butyrate-synthesis pathway of Clostridium acetobutylicum ATCC824." Gene 134: 107-111.

Wang, F. H.;H. J. Qu;D. W. Zhang;P. TianT. W. Tan (2007). "Produciton of 1,3-propanediol from glycerol by recombinant E.coli using incompatible plasmids system." Mol. Biotech. 37(2): 112-119.

Wang, M. Y.;Y. L. Tsai;B. H. OlsonJ. S. Chang (2008). "Monitoring dark hydrogen fermentation performance of indigenous Clostridium butyricum by hydorgenase gene expression using RT-PCR and qPCR." Int. J. Hydrog. Energy 33(18): 4730-4738.

Weber, F. J.A. M. De Bont (1996). "Adaptation mechanisms of microorganisms to the toxic effects of organic solvents on membranes." Biocheimica et Biophysica Acta, Reviews on Biomembranes 1286(3): 225-245.

Weyer, E. R.L. F. Rettger (1927). "A comparative study of six different strains of the organism commonly concerned in large-scale production of butyl alcohol and acetone by the biological process." J. Bacteriol 14: 399-424.

Whang, L. M.;C. J. HsiaoS. S. Cheng (2006). "A dual-substrate steady-state model for biological hydrogen production in an anaerobic hydrogen fermentation

192

process." Biotechnol. Bioeng. 95(3): 492-500.

Williams, E. A.;J. M. CoxheadJ. C. Mathers (2003). "Anti-cancer effects of butyrate: use of microarray technology to investigate mechanisms." Proc. Nutr. Soc. 62: 107-115.

Wisenborn, D. P.;F. B. RudolphE. T. Papoutsakis (1989). "Phosphotransbutyrylase from Clostridium acetobutylicum ATCC824 and its role in acidogenesis." Appl. and Enviro. Microbiol. 66: 317-322.

Woodward, J.;M. Mattingly;M. Danson;D. Hough;N. WardM. Adams (1996). "In vitro hydrogen production by glucose dehydrogenase and hydrogenase." Nat. Biotechnol. 14: 872-874.

Wu, Z. T.S. T. Yang (2003). "Extractive fermentation for butyric acid production from glucose by Clostridium tyrobutyricum." Biotech. Bioeng. 82: 93-102.

Yang, S. T. (1996). "Extractive fermentation using convoluted fibrous bed bioreactor." U.S. patent US 5563069

Yang, S. T.;Y. HuangG. Hong (1995). "A novel recycle batch immobilized cell bioreactor for propionate production from whey lactose." Biotech. Bioeng. 45: 379-386.

Yang, S. T.;H. Zhu;Y. LiG. Hong (1994). "Continuous propionate production from whey permeate using a novel fibrous bed bioreactor." Biotech. Bioeng. 43(11): 1124-1130.

Yat, S. C.;A. BergerD. R. Shonnard (2008). "Kinetic characterization for dilute sulfuric acid hydrolysis of timber varieties and switchgrass." Bioresour Technol 99(9): 3855-63.

Yoneda, N.;S. Kusano;M. Yasui;P. PujadoS. Wilcher (2001). "Recent advances in processes and catalysts for the production of acetic acid." Appl. Cataly. A: Gen. 221: 253-265.

Yoshida, A.;T. Nishimura;H. Kawaguchi;M. InuiH. Yukawa (2005). "Enhanced hydrogen production from formic acid by formate hydrogen lyase-overexpressing Escherichia coli strains." Appl. Environ. Microbiol. 71: 6762-6768.

Yoshida, A.;T. Nishimura;H. Kawaguchi;M. InuiH. Yukawa (2006). "Enhanced hydrogen production from glucose using Idh- and frd-inactivated Escherichia coli strains." Appl. Microbio. Biotech. 73: 67-72.

Yoshida, A.;T. Nishimura;H. Kawaguchi;M. InuiH. Yukawa (2007). "Efficient induction of formate hydrogen lyase of aerobically grown Escherichia coli in a three-step biohydrogen production process." Appl. Microbio. Biotech. 74: 754-760.

193

Yoshimoto, H.;T. Fukushige;T. YonezawaH. Sone (2002). "Genetic and physiological analysis of branched-chain alcohols and isoamyl acetate production in Saccharomyces cerevisiae." Appl. Microbio. Biotech. 59: 501-508.

Young, M.S. Cole (1993). Bacillus subtilis and other gram-positive bacteria: biochemistry, physiology, and molecular genetics. Washington D.C., American Society for Microbiology.

Zaigler, A.;S. C. SchusterJ. Soppa (2003). "Construction and usage of a onefold-coverage shotgun DNA microaaray to characterize the metabolism of the archaeon Haloferax volcanii." Mol. Microbiol. 48: 1089-1105.

Zeng, A. P.;H. Biebl;H. SchliekerW. D. Deckwer (1993). "Pathway analysis of glycerol fermentation by Klebsiella pneumoniae: regulation of reducing equivalent balance and product formation. ." Enzyme Microb. Technol. 15: 770-779.

Zhang, C.;M. Zhang;J. Ju;J. Nietfeldt;J. Wise;P. M. Terry;M. Olson;S. D. Kachman;M. Wiedmann;M. SamadpourA. K. Benson (2003). "Genome diversification in phylogenetic lineages I and II of Listeria monocytogenes: Identification of segments unique to lineage II populations." J. Bacterial 185: 5573-5584.

Zhang, T.;H. LiuH. H. P. Fang (2003). "Biohydrogen production from starch in wastewater under thermophilic conditions. ." J. Environ. Manage. 69: 149-156.

Zhao, Y.;C. A. Tomas;F. B. Rudolph;E. T. PapoutsakisG. N. Bennett (2005). "Intracellular butyryl phosphate and acetyl phosphate concentrations in Clostridium acetobutylicum and their implications for solvent formation." Appl. Environ. Microbiol. 71: 530-537.

Zhu, Y. (2003). Enhanced butyric acid fermentation by Clostridium tyrobutyricum immobilized in a fibrous-bed bioreactor. Department of chemical engineering. Columbus OH, The Ohio State University. Ph.D. dissertation.

Zhu, Y.;X. LiuS. T. Yang (2004). "Construction and characterization of pta gene-deleted mutant of Clostridium tyrobutyricum for enhanced butyric acid fermentation." Biotech. Bioeng. 90: 154-166.

Zhu, Y.;Z. T. WuY. S.T. (2002). "Butyric acid production from acid hydrolysate of corn fiber by Clostridium tyrobutyricum in a fibrous-bed bioreactor." Proc. Biochem. 38(5): 657-666.

Zhu, Y.S. T. Yang (2003). "Adaptation of Clostridium tyrobutylicum for enhanced tolerance to butyric acid in a fibrous-bed bioreactor." Biotech. Prog. 19: 365-372.

194

Zigova, J.E. Sturdisk (2000). "Advances in biotechnological production of butyric acid." J. Indus. Microb. & Biotech. 24(3): 153-160.

Zigova, J.;E. Sturdisk;D. VandakS. Schlosser (1999). "Butyric acid production by Clostridium butyricum with integrated extraction and pertraction." Proc. Biochem. 34: 153-160.

195

APPENDICES

Appendix A Microarray Fabrication

A.1 Genomic library preparation

The chromosome DNA extracted by Qiagen Genomic-tip was partially digested

with Sau3AI. The DNA fragments were separated on 0.7% agarose gel and the fragments

of 1-5 kb were cut, purified and ligated to BamHI digested plasmid pXL. The ligated plasmids were transformed into Max E. coli DH5a competent cells (Invitrogen), and then plated on LB supplemented with 50 μg/ml kanamycin. More than 20,000 white colonies in one standard transformation reaction using E. coli DH5α competent cells have been obtained (Figure A.1).

The white colonies were picked from the agar plates and stored in 96-well plates at -80oC with 15% glycerol for further study. We prepared a 2.5× coverage library which

containing 1920 fragments in 20 plates. The probability of the gene of interest presenting

in the library is 85% according to n = ln(1-p)/ln(1-f), where f = 0.001 (fractional portion

of genome present in a single DNA fragment or the average size of the insert (3 kb)

divided by the genomic size 3 Mb) and n=1920. The cells were harvested, and the

plasmid was extracted by Qiagen spin miniprep kit. After random quality check of the

insert on agarose gel (Figure A.2), the insert genes were amplified by M13 forward

primer (5’-GTAAAACGACGGCCAG-3’) and M13 reverse primer (5’-

196

CAGGAAACAGCTATGAC-3’) and the PCR products were randomly checked (Figure

A.3). The PCR products were purified by isopropanol precipitation, checked by agarose gel again (Figure A.4), and then dissolved in 3XSSC buffer for printing.

A.2 Microarray slide preparation and quality control

The microarray slide printing is done in-house with a high-precision fast motion arrayer with 16 printing heads. For microarray printing, the widely used method is to print the DNA onto poly-lysine coated glass slides. Since the poly-lysine is positively charged, the negatively charged DNA fragment could easily bind to the poly-lysine on the slide surface. The slides are prepared following the standard protocols from Dr.

DeRisi’s Lab in UCSF (http://derisilab.ucsf.edu/pdfs/PolylysineSlides.pdf). Salmon sperm DNA was spotted onto the slides for test printing. SpotQC kit (Integrated DNA

Tech.) was used to hybridize the microarray slides to check the coating and printing quality (Figure A.5).

The random genomic library was printed on the slide with triplicate while controls are located on different parts of the chip. DNA fragments encoding PTA and 16S rRNA are used as positive control, while negative control is salmon sperm DNA which shows no homologies to the C. tyrobutyricum.

After spotting, UV cross-linking was used to immobilize the DNA fragments on the slide followed by post-processing which consists of re-hydration and blocking. Re- hydration allows the DNA to evenly distribute across the surface of the spot to eliminate the donut-shaped appearance of the spots and to increase the amount of total DNA bound.

197

Blocking is required to prevent the non-specific binding between the labeled sample and

the exposed amine on the slide surface. Succinic anhydride reacts with amine and has the

ability to cap the free amines.

Because all of the DNA fragments are PCR products amplified by M13 primers, the primers sequence exist in each spot with similar amount. The cy3 labeled M13 primer

will be used to check the consistency of the slide as a quality control method after each batch of printing.

A.3 RNA extraction and cDNA preparation, dye labeling, and competitive hybridization

Standard procedures were followed to extract total RNA from C. tyrobutyricum using trizol reagent. Briefly, cells cultured under different conditions were harvested and the total RNA will be extracted using Trizol reagent (Invitrogen) following the manufacture’s standard protocol. The RNA concentration was be determined by spectrometry and the integrity was checked by agarose gel electrophoresis. Indirect labeling was employed by which aminoally-dUTP is firstly incorporated into the cDNA during reverse transcription, and then the dye is coupled to the function group amino-ally.

The cDNA was synthesized using the tRNA samples from control and test sample (wild type and mutant) as the template. The reverse transcription was performed with (SuperScript III, Invitrogen), random hexamers and dNTP mixture containing 10% aa-dUTP. The cy3 and cy5 dyes was incubated with control and test

198

samples for labeling and the labeled cDNA will be pooled together for the following competitive hybridization.

Printed microarray slide was washed to remove unbound materials and the double-stranded DNA will be denatured by boiling. The microarray slide will be competitively hybridized in hybridization cassettes (Biorad) for 5 h at 42°C under a flat cover slips with 5.0 µl probe solution containing 5×SSC + 0.2% SDS + cye-labeled cDNA sample pool. Hybridized microarray will be washed twice for 5 min each in 2×

SSC + 0.2% SDS at 25°C, once for 1 min in 2× SSC at 25°C and spun dry for 1 min at

500 ×g. The hybridization condition needs to be optimized to achieve satisfying results.

The duplicated experiment will be performed.

The mRNA purification is an option to remove the ribosome RNA which counts for more than 90% of total RNA if the differentiation hybridization result is too weak to draw any conclusion.

A.4 Scanning and data analysis

Hybridized microarray will be scanned by a multi-wavelength laser microscope.

The resulting fluorescence is measured at 570nm and 670 nm after excited at 550nm and

649nm for cy3 and cy5 respectively. The Cy3 and Cy5 signals will be quantified and the average of the measurements will be calculated. We will scatter plot the signal to examine the gene expression level. Most signals are supposed to be near the diagonal because of the constitutive expression. The gene fragments which have different transcript level will be spotted. The data from these spots will be clustered using the

199 standard correlation, and different clusters of co-regulated genes will be detected.

Furthermore, these gene fragments will be sequenced, and identified by comparison with the 454 database.

The whole process described above is illustrated by Figure A.6.

200

Figure A.1 White/blue screening for gene library construction

201

Figure A.2 Random quality check of plasmids.

Figure A.3 Random quality check of PCR product.

Figure A.4 Random quality check of purification.

202

Figure A.5 Photograph of printed microarray slide and the scanned results after hybridization with cy3-labeled oligomers.

203

A. Gene library construction B. Microarray chip construction C. Hybridization and analysis

C. tyrobutyricum cell Chromosomal DNA isolation Sample mRNA extraction Reference Insert amplification by PCR Colony Chromosomal DNA Reverse transcription (partial) digestion mRNA mRNA mRNA labeling

DNA fragment (insert) Insert Insert purification Ligation

Cloning vector Green Cy3 Red Cy5 Plasmids Array printing Probe Transformation Competitive E. coli DH5α hybridization

White/blue screen Arrayer Colony transfer to 96 format Store arrayer slides Scan and analyze

Store 96 well plates

Sequence 2D/MS G A T C Analyze Slide storage cassette Gene library

Figure A.6 Scheme of shotgun DNA microarray construction. A. Construction of gene library; B. Construction of microarray gene chips; C. RNA labeling, chip hybridization, and data analysis. (adapted from Liu, 2005)

204

Appendix B Summary of CoA transferase and Genes involved in glycolysis.

The 454 sequencing produced 211,099 good reads, containing 55,970,852 bp of raw sequence. Assembly produced 69 contigs, 3,016,691 bp.

The genes encoding the enzymes involved in the metabolic pathway from pyruvate to various end products, including acidogenesis and solventogenesis according to ATCC 824, found and annotated in the 454 database are summarized in Table 5.2.

Other genes responsible for enzymes in glycolysis are also summarized in Figure B.1 and

Table B.1. Furthermore, two genes encoding for acetoacetyl-CoA: acetate/butyrate CoA transferase (EC# 2.8.3.9) are annotated in the database, one of them is located in the contig 64 starting at 129 bp and ending at 1682 bp and the other one is from 21339 bp to

22931 bp in the contig 22. The deduced amino acid sequences from these two genes show

58.4 % identity to each other. The one located in contig 64 shows 80% identity to the

CoA transferase from Clostridium Novyi (Gi: 4541513) and the other one shows 76% to the same protein. The consensus against other bacterial CoA transferases is summarized in Table B.2.

205

Figure B.1 Genes annotated in the 454 sequencing database for the glycolysis pathway of C. tyrobutyricum.

206

# Identity: 310/531 (58.4%) # Similarity: 405/531 (76.3%) # Gaps: 15/531 ( 2.8%) # Score: 1648.0

CTF2 1 MGVKFLKACDAVEMVKDGDLIATSGFVGSLCSEELSIALEKRFEETGHPK 50 |..|.:....||:::||||.:|..||||....|:::..:|:.:.|...|| CTF1 1 MKSKLMSKDKAVKLIKDGDTVAVGGFVGCAHPEDITSEIEENYIENRTPK 50

CTF2 51 NLTLVYSAAQGDGKGKGADHFAHEGMLKRVAGGHWNLSPKLGKMAVENKI 100 ||||:|:|.|||...:|.:||.|||::.:|.||||.|||||.|:|::||| CTF1 51 NLTLIYAAGQGDSADRGLNHFGHEGLVSKVIGGHWALSPKLQKLALDNKI 100

CTF2 101 EAYNLPQGTIAQLFRDIAGKRIGTITHVGLNTFVDPRIQGGKLNDITKED 150 ||||||||.|:||:||||.||.||||||||.||:|||::|||||.||||: CTF1 101 EAYNLPQGIISQLYRDIAAKRPGTITHVGLKTFIDPRLEGGKLNTITKEN 150

CTF2 151 VVKLINIEGQEKLLYKAFPIDVCFLRGSFADENGNISLEREVGTIDVLSI 200 :|:||||.|:|.|.||..|:||..||.::|||.||.::|:|...:|..:: CTF1 151 IVELINIGGKEYLFYKTIPLDVVILRATYADEFGNATMEKEAAILDATAM 200

CTF2 201 AQACRNSGGIVIVQVEKVVSSGTLDPRLVKIPGIYVDVVVVAKPEHHEQF 250 |||.:|||||||||||:|||.|:|||:.|||||||||.:||::|::|.|. CTF1 201 AQAAKNSGGIVIVQVEQVVSKGSLDPKKVKIPGIYVDAIVVSQPKNHMQT 250

CTF2 251 FGCNGYDPALTGEVRVPLKAIQYPELNARKIIARRAAMELKPDAVVNLGI 300 |..| |:|:.:||.|..:.:|...:||.||:|||||||||.|::|.|||| CTF1 251 FSEN-YNPSYSGEARFLVNSIVPMQLNERKVIARRAAMELVPNSVTNLGI 299

CTF2 301 GIPEVISMVANEEEIGEYMTMTVEAGAVGGVPLGGTGFGACINPESILDE 350 ||||.|:.|||||.|.:.||:|:|:|.:||||.||..|||..||:||||: CTF1 300 GIPEGIATVANEEGIADEMTLTIESGGIGGVPSGGLSFGASTNPQSILDQ 349

CTF2 351 AYQFDFYDGGGTDIAFLGLAQVDRHGNINVSKFGPRIAGCGGFINITQNA 400 |.|||:|||||.|:|||||||.||.||||||||||:||||||||||:||: CTF1 350 ASQFDYYDGGGLDVAFLGLAQCDRDGNINVSKFGPKIAGCGGFINISQNS 399

CTF2 401 KRVLFCGTFTAGGLKIDTGNGKLKIIQEGRSIKFLNEVEQITFSGEYAIK 450 |:|::||||||||||:...||||.|.::|:..||::.||||:|||:||.. CTF1 400 KKVVYCGTFTAGGLKVKVENGKLNIEKDGKFNKFIDTVEQISFSGQYAQS 449

CTF2 451 TKQPVTYITERAVFELKEEGLYITEIAPGIDLEKDVLGLMEFKPKIDKSL 500 ..|.|.||||||||.|.:|||::.|||||||::||:|..|:|:|||.::| CTF1 450 IGQTVLYITERAVFRLTKEGLFLEEIAPGIDMKKDILDHMDFRPKISENL 499

CTF2 501 KLMDEKIFHNGLMGLKK 517 |:|||:||....:|:.. CTF1 500 KIMDERIFKEEPIGINIGQSSKEFPGELSNV 530

Figure B.2 Alignment of two CoA transferases in C. tyrobutyricum

207

Enzyme Name GLUCOSE/XYLOSE TO PYRUVATE contig Name Glycogen Synthase 1 CDS_8 Triose Phosphate 3 tpiA Triose Phosphate Isomerase 13 tpiA

Glucose-6-Phosphate Isomerase (Phosphoglucose isomerase) 4 pgi Fructose 1,6 bisphosphate aldolase class II 13 fba fructose-bisphosphate aldolase 44 fbaA Fructose 1,6 bisphosphatase 35 CDS_20 Glyceraldehyde 3-phosphate dehydrogenase,type 1 13 gap 13 pgk 2,3-bisphosphoglycerate-independent phosphoglycerate mutase 13 gpmI phosphoglycerate mutase (possible) 19 CDS_30 phosphoglycerate mutase (possible) 19 CDS_31 phosphopyruvate hydratase (enolase) 13 eno 1 13 CDS_77 18 glcK phospho-2-dehydro- 3-deoxyheptonate aldolase 24 aroF phosphoglucomutase 25 glmM nucleoside-diphosphate sugar pyrophosphorylase(ADP- glucose pyrophosphorylase) 28 gcd Pyruvate dehydrogenase complex component 3(E2) 28 CDS_40 Pyruvate dehydrogenase complex (E3) 28 lpdA Pyruvate dehydrogenase complex (E3) (NADH oxidase) 28 CDS_142 6- 34 pfkA 6-phosphofructokinase 58 pfkA Xylose isomerase 33 xylA Xylulokinase 33 xylB 58 pyk transketolase C-terminal section 34 CDS_81 transketolase N-terminal section 34 tktN 1- deoxy D-xylulose 5 phosphate synthase 36 dxs phosphoenolpyruvate protein phosphotransferase 40 ptsP pyruvate carboxyltransferase 41 CDS_4 1-phosphofructokinase 44 pfkB

Table B.1 Location of genes involved in glycolysis in the 454 sequencing database of C. tyrobutyricum.

208

CoA transferase on contig CoA transferase on contig 64 22

strain identity% similiarity% identity% similiarity%

C. beijerinckii 72 87 76 89

C. acetobutylicum 26 44 27 42

C. novyi 64 80 58 76

C. perferingens 75 86 59 77

C. tetani 74 87 65 82

C. butylicum 54 70 52 70

C. difficile 27 45 30 50

C. botulicum 73 87 74 87

Table B.2 Comparison of amino acids sequences of two CoA transferases from C. tyrobutyricum against other bacteria.

209

Appendix C Chromatography methods and standard spectrum

C.1 High performance liquid chromatography

The high performance liquid chromatography (HPLC) system consisted of an automatic injector (Shimadzu SIL-10Ai), a pump (Shimadzu LC-10Ai), an organic acid analysis column (Bio-Rad HPX-87H), a column oven at 45°C (Shimadzu CTO-10A), and a reflective index detector (Shimadzu RID-10A). The eluent was 0.01 N H2SO4 at a flow rate of 0.6 ml/min. Samples, each with a volume of 15 µl, were injected by the auto- sampler. The running time for each sample was 25 min. Peak height was used to calculate the concentration of each component based on the analysis of the standard mixture containing all the compounds at 2 g/l. Samples of cell-free fermentation broth were diluted with distilled water at different ratios from 1/6 to 1/20 depending on the concentration of the compounds to be analyzed. A standard chromatogram is shown in

Figure C.1.

C.2 Gas chromatography for gas production

An on-line gas chromatograph (Shimadzu GC 2014) was used for automatic measurement of gases (H2 and CO2) produced during the fermentation of C. tyrobutyricum. The fermenter was connected to GC and 100 µl of gas was collected and injected by an automatic six-head valve to analyze the composition of gas produced in the fermentation with a thermo conductivity detector and RT-QPLOT column (Restek).

210

Argon was used as the carrier gas at flow rate of 4.38 ml/min. The injector, column, and detector temperatures were maintained at 60°C, 30°C, and 60°C, respectively. A standard chromatogram is shown in Figure C.2.

C.3 Gas chromatography for alcohol analysis

The composition of alcohol produced in the fermentation was analyzed with a gas

chromatograph (Shimadzu GC 2014) equipped with Flame Ionization Detector and

Stabilwax-DA column (Restek) with helium as the carrier gas at the flow rate of 2

ml/min. 10 µl of sample in concentration range between 0.125 g/l to 2g/l was injected for the analysis. The injector temperature and the detector temperature were set to 200°C.

The column temperature was programmed as the follows: 80°C hold for 3 min, and then increased to 150 °C at 30°C/min and hold at 150°C for 4 min. A standard chromatogram is shown in Figure C.3.

211

50 50 RID-10A 052709 Name Retention Time Height 45 45 Area ESTD concentration

40 40

35 35

30 30

25 25 lactose 7.467 21560 324634 2.000 7.467 lactose glucose 8.717 21230 327458 2.000 8.717 glucose xylose 9.333 19940 318552 2.000 19940 318552 9.333 xylose 20 20 uRIU uRIU

15 15

10 8898 169126 2.000 12.317 acid lactic 10 acetic acid 14.567 7513 151929 2.000 7513 151929 14.567 acid acetic butyric acid 21.083 7197 215982 2.000 21.083 acid butyric

5 5 11.617 1542 35073 0.000 11.100 1279 24650 0.000 1279 11.100

0 0

-5 -5

-10 -10 0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 Minutes

Figure C.1 HPLC chromatogram for standard sample containing glucose, xylose, lactose, acetic acid, lactic acid and butyric acid (2 g/l each).

212

uV(x10,000) 2.50 Chromatogram

2.25

2.00

1.75 H2/5.119

1.50

1.25

1.00

0.75

0.50

0.25 CO2/7.604 0.00

-0.25 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0 11.0 min

Figure C.2 GC chromatogram for the standard sample containing hydrogen (80 %) and

CO2 (20 %).

213

uV(x10,000) 5.0 Chromatogram

4.5

4.0 Acetic Acid/6.132 Acetic Butyric Acid/7.693 Butyric Acetone/1.533 Butanol/3.573 3.5

3.0

2.5 Ethanol/1.896

2.0

1.5

1.0

0.5 /4.225 /6.818 Acetic Acid/6.314 Acetic Acetic Acid/5.869 Acetic Butyric Acid/7.340 Butyric /5.613 Ethanol/1.762 /8.202 /5.449 /7.055 /5.335 /4.917 /5.183 /4.401 /1.402 0.0 /1.294

-0.5 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 min

Figure C.3 GC chromatogram for standard sample containing acetone, ethanol, butanol,

acetic acid and butyric acid (0.5 g/l each).

214