ENGINEERING ACTINOBACILLUS SUCCINOGENES FOR SUCCINATE PRODUCTION- A FOCUS ON SUCCINATE TRANSPORTERS AND SMALL RNAs By Rajasi Virendra Joshi A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Microbiology and Molecular Genetics- Doctor of Philosophy 2017 ABSTRACT ENGINEERING ACTINOBACILLUS SUCCINOGENES FOR SUCCINATE PRODUCTION- A FOCUS ON SUCCINATE TRANSPORTERS AND SMALL RNAS By Rajasi Virendra Joshi An important aspect of any industrial scale bio-based production is the choice of biocatalyst used. Many commercially relevant microorganisms and industrial strains have been engineered to optimize the production of bio-based chemicals. One such chemical is succinate, listed as one of the top 12 building block chemicals from biomass by the US Department of Energy. Succinate is considered an important platform chemical as it has a number of applications and, most importantly, is a precursor to high-volume value-added commodity chemicals. Bio-based succinate is currently being produced at industrial scale levels using engineered microorganisms such as E. coli and S. cerevisiae. Actinobacillus succinogenes is one of the best natural succinate producers, which can grow on a wide variety of substrates, and, with the advance in genetic tools, can possibly be engineered for increased succinate production. Very few studies have focused on using succinate exporters as metabolic engineering targets for succinate production. Only a handful of studies have been carried out in E. coli and C. glutamicum with none whatsoever in A. succinogenes. With a combination of proteomics and transcriptomics we have identified candidate succinate transporters in A. succinogenes. Four of the top hits in our proteomics analysis were Asuc_1999, Asuc_0142, Asuc_2058 and Asuc_1990-91. To carefully tune the expression of these membrane proteins, we generated a library of promoters covering a large range of strengths below the strong, constitutive promoter (ppckA) we had been using. Some of the promoters were truncated versions of ppckA, and others were identified from our transcriptomics data. These promoters were tested using lacZ as the reporter gene in an A. succinogenes ∆lacZ background. Promoters ranged from ppckA as the highest down to pAsuc_0701 with a strength 209-fold lower than ppckA. The four succinate transporter candidates over-expressed under ppckA-92, a truncated version of ppckA, increased the succinate yield in glucose cultures compared to the control strain carrying the empty vector. Synthetic small RNAs (sRNAs) are another tool for metabolically engineering industrially relevant microorganisms. However, no sRNAs have been identified in A. succinogenes and only a few have been identified in other members of the Pasteurellaceae family. We identified sRNAs in A. succinogenes grown anaerobically on glucose and microaerobically on glycerol by RNA sequencing. We found 260 sRNAs in total, of which 39 were predicted by at least one of five computational programs. We validated 14 sRNAs identified from sequencing with RT-PCR. Additionally we identified probable Hfq-binding sRNAs through their Rho-independent terminators, a key feature of Hfq-binding sRNAs. Using additional characteristics of Hfqbinding sRNAs, we designed synthetic sRNAs targeting lacZ expression as a proof of concept. One plasmid-borne synthetic sRNA caused a 32% decrease in β-galactosidase activity in lactosegrown cultures. Using the same sRNAs scaffolds, we generated synthetic sRNAs that targeted the ackA and pta mRNAs to decrease the production of acetate, one of the major by-products of succinate production. One of the synthetic sRNAs targeting ackA caused a 14% decrease in the acetate yield of glucose-grown cultures. In summary, we have identified candidate succinate transporters and seen an increase in succinate production upon their overexpression. In the process, we have developed a promoter library for tunable expression of genes in A. succinogenes. We have also shown that sRNAs can be used as a tool for metabolic engineering in A. succinogenes, although additional studies are needed to make it more tunable and robust. This dissertation is dedicated to my entire family, especially my husband, Ashwin for his patience, constant encouragement and morale boosting pep talks. I thank my parents, sister, brother-in law and in-laws for always believing in me and supporting me throughout this journey. iv ACKNOWLEDGEMENTS This dissertation has been possible only because of the contributions of several people who have guided and helped me throughout these years. First and foremost, I would like to express my sincere gratitude to my mentor, Dr. Claire Vieille for giving me the opportunity to work in her lab during my Master’s and for my PhD. I thank her for being a great mentor and for always taking out the time to discuss any idea, big or small, despite being busy with numerous other things. Her intellect, passion for science and positive attitude has always inspired and motivated me, especially during the more trying times of this project. I will always be indebted to her, for her guidance and thoughtful input throughout my PhD. I am thankful to all the members of the Vieille lab- Maris Laiveneiks, Dr. Bryan Schindler, Dr. Justin Beauchamp, Dr. Nikolas Mcpherson and Peggy Wolf for a great work atmosphere with lots of support and help. I am truly grateful for the time taken by Dr. Bryan Schindler to train me during my initial years in the lab even when he was busy with his own experiments. I am also grateful to Maris Laivenieks, who was always willing to help me, especially during my initial struggles while building a chemostat and mixing gas tanks. I would like to thank Dr. Nikolas McPherson for being an excellent lab member and for all the helpful discussions we had while working with Actinobacillus succinogenes during the course of both of our PhDs. I must acknowledge Dr. Justin Beauchamp for helping me with the intial RNAseq analysis. I wish to thank Barry Tigner, who has saved my experiments numerous times, by fixing last minute instrument fails. I must thank my committee members, Drs. Gemma Reguera, Michael Bagdasarian, Shannon Manning, Yair Shachar-Hill for their expert guidance and constructive feedback. I v would like to thank Dr. Donna Koslowsky for letting me use her lab and Laura Kirby, her grad student for helping me set up Northern blots. I would like to thank all the undergraduates who helped me with my project, Alex Yang, Jean Kim, Maeva Bottex, Matthias Schulz, Kanupriya Tiwari, Joel Tencer, Jacob Gibson, Aaron Walters, Jack Peleman and Corey Kennelly. It was a pleasure interacting with each of them and training them has made me a better researcher. I would like to thank Betty Miller, graduate secretary of Microbiology and Molecular Genetics for helping me navigate through the paperwork required and other assistance throughout these years. I am grateful for the lasting friendships I have made here at MSU. I feel fortunate to have had a strong support system in them, sharing the ups and downs of graduate life. vi TABLE OF CONTENTS LIST OF TABLES .......................................................................................................................... x LIST OF FIGURES ...................................................................................................................... xii Chapter 1 Introduction .................................................................................................................... 1 1.1 Bio-based succinate production ............................................................................................ 2 1.2 Substrates and feedstocks for succinate production .............................................................. 4 1.3 Applications for succinate and its derivatives ....................................................................... 4 1.4 Succinate-producing microorganisms ................................................................................... 5 1.4.1 Native succinate producers ............................................................................................. 5 1.4.2 Engineered succinate producers ................................................................................... 12 1.5 C4-dicarboxylate transporters ............................................................................................. 21 1.5.1 E. coli C4-dicarboxylate transporters ........................................................................... 21 1.5.2 C4-dicarboxylate transporters in other succinate producers ........................................ 23 1.6 sRNAs as synthetic regulators for metabolic engineering .................................................. 24 1.7 Identification of small RNAs in Pasteurelleaceae ............................................................... 28 1.8 Introduction to chapters ....................................................................................................... 28 REFERENCES .......................................................................................................................... 31 Chapter 2 Identification of succinate transporters in Actinobacillus succinogenes ...................... 42 2.1 Abstract ............................................................................................................................... 43 2.2 Introduction ......................................................................................................................... 44 2.3 Materials and Methods ........................................................................................................ 47 2.3.1 Strains, media, and culture conditions. ......................................................................... 47 2.3.2 Plasmids, DNA manipulations, and electroporations. .................................................. 47 2.3.3 Preparation of A. succinogenes cytoplasmic membrane-enriched fractions for proteomics analysis. .............................................................................................................. 49 2.3.4 Proteomic analysis. ....................................................................................................... 50 2.3.5 Total RNA purification and RNA sequencing ............................................................. 52 2.3.6 Identification of A. succinogenes’s pckA transcription start site. ................................. 53 2.3.7 Construction of truncated pckA promoters. .................................................................. 53 2.3.8 Construction of an A. succinogenes ΔlacZ mutant strain. ............................................ 54 2.3.9 Preparation of crude extracts, β-galactosidase assays, and protein assays. .................. 55 2.3.10 Construction of a promoter library based on RNAseq results. ................................... 55 2.3.11 Construction of strains expressing putative succinate transporters. ........................... 56 2.3.12 High-performance liquid chromatography (HPLC) analysis of fermentation media from strains expressing transporters. ..................................................................................... 57 2.4 Results and Discussion ........................................................................................................ 57 2.4.1 Potential C4-dicarboxylate exporters in A. succinogenes ............................................. 57 2.4.2 Proteomic analysis of A. succinogenes cytoplasmic membrane-enriched fractions .... 59 2.4.3 Transcript levels of A. succinogenes putative C4-dicarboxylate transporters .............. 62 2.4.4 Construction of an expression reporter system. ........................................................... 64 vii 2.4.5 Identification of ppckA’s TSS by RLM-RACE and construction of truncated ppckA promoters ............................................................................................................................... 65 2.4.6 Identification of promoters of different strengths from the transcriptomics results. .... 67 2.4.7 Overexpression of putative succinate transporter candidates in A. succinogenes and succinate production. ............................................................................................................. 68 2.4.8 The A. succinogenes succinate efflux transporters differ from those of E. coli and C. glutamicum ............................................................................................................................ 71 2.5 Conclusion........................................................................................................................... 73 2.6 Acknowledgements ............................................................................................................. 73 APPENDIX ............................................................................................................................... 74 REFERENCES .......................................................................................................................... 78 Chapter 3 Identification of sRNA in Actinobacillus succinogenes 130Z ..................................... 82 3.1 Abstract ............................................................................................................................... 83 3.1 Introduction ......................................................................................................................... 84 3.2 Materials and Methods ........................................................................................................ 87 3.2.1 Bacterial strains and growth conditions. ...................................................................... 87 3.2.2 RNA isolation, library preparation, and sequencing. ................................................... 89 3.2.3 Trimming of reads, mapping, and sRNA identification. .............................................. 89 3.2.4 Small RNA analysis. .................................................................................................... 90 3.2.5 Computational prediction of sRNAs. ........................................................................... 91 3.2.6 RT-PCR validation of small RNAs. ............................................................................. 92 3.2.7 Construction of synthetic small RNA constructs. ........................................................ 92 3.2.8 Beta-galactosidase assays. ............................................................................................ 93 3.2.9 Determination of fermentation balances. ..................................................................... 94 3.3 Results and Discussion ........................................................................................................ 95 3.3.1 Detection of sRNAs from RNAseq data using RockHopper. ...................................... 95 3.3.2 Comparison of RNAseq-identified to computationally predicted sRNAs. ................ 104 3.3.3 sRNAs with Rho-independent terminators. ................................................................ 106 3.3.4 RT-PCR validation of small RNAs. ........................................................................... 106 3.3.5 Scaffold selection for synthetic sRNA design based on predicted Hfq-dependent small RNAs. .................................................................................................................................. 110 3.3.6 Testing of synthetic sRNA designs with lacZ as target gene. .................................... 111 3.3.7 Synthetic sRNAs for inhibiting acetate production. ................................................... 114 3.4 Conclusion......................................................................................................................... 117 3.5 Acknowledgements ........................................................................................................... 117 APPENDIX ............................................................................................................................. 119 REFERENCES ........................................................................................................................ 131 Chapter 4 Development of a markerless knockout method for Actinobacillus succinogenes .... 137 4.1 Abstract ............................................................................................................................. 140 4.2 Introduction ....................................................................................................................... 141 4.3 Materials and Methods ...................................................................................................... 144 4.3.1 Strains, media, culture conditions, and chemicals. ..................................................... 144 4.3.2 Plasmids, DNA manipulations, and electroporations. ................................................ 146 4.3.3 Construction of plasmid pLGZ924. ............................................................................ 147 viii 4.3.4 Construction of Δfrd::icd and ΔpflB::icd mutants. ..................................................... 148 4.3.5 Construction of the ∆pflB ΔlacZ::icd, ∆pflB Δcit::icd, and ∆pflB Δacn::icd double mutants. ............................................................................................................................... 149 4.3.6 Natural transformation................................................................................................ 150 4.3.7 Removal of the icd marker. ........................................................................................ 151 4.3.8 Curing of plasmid pCV933. ....................................................................................... 151 4.3.9 Enzyme assays. ........................................................................................................... 152 4.3.10 Analysis of fermentation products............................................................................ 153 4.4 Results ............................................................................................................................... 153 4.4.1 A positive selection method for recombination events in A. succinogenes. .............. 153 4.4.2 Natural transformation of A. succinogenes using a Δfrd::icd construct. .................... 154 4.4.3 Confirmation of the Δfrd::icd knockout strain. .......................................................... 155 4.4.4 Excision of the selection marker. ............................................................................... 156 4.4.5 Curing pCV933 from A. succinogenes Δfrd(pCV933). ............................................. 156 4.4.6 Construction of strain ΔpflB. ...................................................................................... 157 4.4.7 Effects of the DNA construct length and of the length of the homologous regions on the efficiency of homologous recombination. ..................................................................... 159 4.4.8 Construction of double knockout mutants. ................................................................. 161 4.4.9 Development of a selective solid medium.................................................................. 163 4.4.10 Knockout mutants via electroporation...................................................................... 164 4.5 Discussion ......................................................................................................................... 165 4.6 Acknowledgements ........................................................................................................... 167 REFERENCES ........................................................................................................................ 172 Chapter 5 Respiratory glycerol metabolism of Actinobacillus succinogenes 130Z for succinate production ................................................................................................................................... 176 Chapter 6 Conclusions and future directions .............................................................................. 178 6.1 Introduction ....................................................................................................................... 179 6.2 Succinate transporters in A. succinogenes......................................................................... 179 6.3 Identification of sRNAs in A. succinogenes...................................................................... 180 6.4 Future directions ................................................................................................................ 181 6.4.1 Experimental identification of Hfq-binding sRNAs................................................... 181 6.4.2 Continuing the development of synthetic sRNAs in A. succinogenes ....................... 181 6.4.3 Identifying the major succinate transporter(s) in A. succinogenes ............................. 182 6.4.4 Construction of a ∆ackA knockout mutant ..................................................................... 182 6.4.5 Combination of different approaches to increase succinate production ..................... 183 REFERENCES ........................................................................................................................... 184 ix LIST OF TABLES Table 1.1 Succinate-producing strains with their yields, productivities, and titers ...................... 20 Table 2.1 Strains and plasmids used in this study ........................................................................ 48 Table 2.2 A. succinogenes putative C4-dicarboxylate transporters .............................................. 58 Table 2.3 Proteomics and transcriptomics data for succinate transporters in A. succinogenes .... 61 Table 2.4 Fermentation balances of strain 130Z over-expressing candidate succinate transporters. ....................................................................................................................................................... 70 Table A2.1 Primers used in this study .......................................................................................... 75 Table A2.2 Transcript levels of 40 candidate genes with low variation across several conditions, compared to pckA .......................................................................................................................... 77 Table 3.1 Strains and plasmids used in this study ........................................................................ 88 Table 3.2 Statistics on read alignments by Rockhopper ............................................................... 96 Table 3.3 A. succinogenes sRNAs detected by RNAseq and Rockhopper analysis in anaerobically-grown glucose cultures and microaerobically-grown glycerol cultures ................ 98 Table 3.4 Rho-independent smRNAs predicted by ARNold ...................................................... 107 Table 3.5 smRNA sequences that fit the criteria for Hfq-dependent sRNAs ............................. 111 Table 3.6 Fermentation balances of strain 130Z expressing synthetic sRNA constructs. .......... 116 Table A3.1 Manually identified smRNAs using alignment files created by Rockhopper.......... 120 Table A3.2 smRNAs common between SIPHT and RNAseq .................................................... 124 Table A3.3 smRNAs common between BSRD and RNAseq..................................................... 125 Table A3.4 smRNAs common between INFERNAL and RNAseq ........................................... 126 Table A3.5 smRNAs common between RNAz and RNAseq ..................................................... 127 Table A3.6 smRNAs common between Rfam and RNAseq ...................................................... 128 Table A3.7 Primers used for RT-PCR validation of sRNAs ...................................................... 129 x Table A3.8 Primers and g-blocks used in this study ................................................................... 130 Table 4.1 Strains and plasmids used in this study ...................................................................... 145 Table 4.2 Frequencies of knockout mutations introduced by natural transformation in A. succinogenes strains 130Z and ΔpflB ......................................................................................... 162 Table A4.1 Oligonucleotide primers used in this study .............................................................. 169 xi LIST OF FIGURES Figure 1.1 Routes to succinate production ...................................................................................... 3 Figure 1.2 Metabolic map of Actinobacillus succinogenes. ......................................................... 16 Figure 1.3 Pathways to succinate production by mixed acid metabolism in bacteria .................. 19 Figure 1.4 Hfq-dependent sRNA organization and structure in E. coli. ....................................... 27 Figure 2.1 Construction of the ΔlacZ strain .................................................................................. 64 Figure 2.2 Identification of the pckA transcription start site. ........................................................ 65 Figure 2.3 Reporter constructs with lacZ under control of truncated versions of pckA’s 5’UTR. 66 Figure 2.4 β-galactosidase activity of promoter reporter constructs............................................. 68 Figure 3.1 Flow chart of the approach used for sRNA identification in A. succinogenes. ........... 95 Figure 3.2 Classification of identified sRNAs. ............................................................................. 97 Figure 3.3 Example of 3’UTR sRNA overlapping with the mRNA transcript and downstream intergenic region with both mRNA and sRNA encoded on the reverse strand. ......................... 103 Figure 3.4 smRNA76 is seen as a single transcript in IGV. ....................................................... 105 Figure 3.5 RT-PCR validation of small RNAs. .......................................................................... 109 Figure 3.6 A. succinogenes smRNA8’s genomic locus .............................................................. 111 Figure 3.7 Synthetic sRNAs targeting lacZ expression. ............................................................. 113 Figure 3.8 β-galactosidase activity of strain 130Z expressing synthetic sRNAs targeting lacZ expression. .................................................................................................................................. 113 Figure 3.9 Synthetic sRNA designs targeting expression of ackA and pta. ............................... 115 Figure 4.1 Construction of the icd selection cassette and construction of strain ∆frd. ............... 148 Figure 4.2 Physical maps of pCR2.1-ΔpflB::icd and its truncated constructs ........................... 157 Figure 4.3 Construction of strain ∆pflB (A) and natural transformation with the truncated products ΔpflB::icd-600, ΔpflB::icd-400, and ΔpflB::icd-200 (B). ............................................. 158 xii Figure 4.4 UV spectra of the HPLC profiles of fermentation supernatants of A. succinogenes 130Z, and ΔpflB grown on AM3 in the presence of 25 mM NaHCO3. ..................................... 159 Figure 4.5 Phenotypic characterization of double knockout mutant ∆pflB-∆lacZ::icd. ............. 163 Figure A4.1 Verification of the construction of strains ∆pflB ∆cit and ∆pflB ∆acn by PCR. .... 171 xiii Chapter 1 Introduction 1 1.1 Bio-based succinate production A huge push towards bio-based succinate production has taken place in the last ten years. Succinate, a C4-dicarboxylic acid, has consistently been among the top ten bio-based chemicals since 2004 (1-3). Succinate is currently produced by two methods. The majority of succinate is produced by catalytic hydrogenation of petroleum-based maleic anhydride, but bio-based succinate production by fermentation of glucose by natural/engineered succinate producers is also catching up, as a number of companies are either in the process of setting up or have already started large-scale fermentation facilities for succinic acid. The global production rate for succinate is between 30,000-50,000 metric tons per year. The market volume for succinate in 2011 was 40,000 metric tons, out of which only 1,150 metric tons were bio-based succinate (4). Market price for petroleum-derived succinate is between $6,000 and $9,000/metric ton (5). If all announced large-scale fermentation projects come to fruition, an estimated total of 140,000 metric tons per year of bio-based succinate will be available globally (5). Large-scale production facilities are estimated to produce a minimum of 642,450 metric tons in 2020. As of today, four major companies have commercial-scale production facilities for bio-based succinic acid—BioAmber, Myriant, Succinity GmbH, and Reverdia. BioAmber is a partnership between DNP Green Technology and the French agricultural cooperative. In 2008, they started a succinic acid plant in Pomacle, France, with an annual capacity of 2,000 metric tons (Chemicals from Biomass) using Escherichia coli (licensed DuPont technology). In August 2015, BioAmber announced the opening of the world’s largest succinate plant in Sarnia, Canada, as a joint venture with Mitsui & Co., Ltd. The company has hit its target for operational milestones, which were laid out at the start, and sold succinic acid worth $3.7 million in the third quarter of 2016 2 (Michael McCoy, ACS). BioAmber has two additional production plants planned or underway with an annual capacity of 65,000 metric tons in USA or Brazil and in Thailand (5). Figure 1.1 Routes to succinate production Myriant started its flagship production facility in Lake Providence, Louisiana, in 2013 with a 13,607 metric tons succinate production capacity per year. It had plans to expand this production facility to have an additional 63,502 metric tons capacity per year (6) by 2015. In partnership with ThyssenKrupp Uhde, Myriant also started a production facility in Leuna, Germany, with a 1,500 metric tons production capacity per year (6). A jointly owned facility in Nanjing, China, between Myriant and China Nation BlueStar is being planned with a potential to scale up capacity to 100,000 metric tons per year (5). Succinity GmbH is a joint venture between BASF and Corbian Purac. BASF has patents on a genetically engineered strain of Basfia succiniciproducens to produce succinate from either glycerol or mixed sugars while fixing the greenhouse gas CO2. A production facility in Montmelό, Spain, has the capacity to produce 10,000 metric tons of succinate per year using crude glycerol as the feedstock. A larger 50,000 metric tons per year facility was planned but no specifics are known thus far (5). 3 Reverdia is a joint venture between Royal DSM, Netherlands, and Roquette Frѐres, France. In 2012, Reverdia started a production facility to produce 10,000 metric tons of succinate per year in Cassano Spinola, Italy, which uses the yeast Saccharomyces cerevisiae to convert commodity sugars to succinate (5). 1.2 Substrates and feedstocks for succinate production Substrates and feedstocks used in fermentations for succinate production can significantly contribute to production costs and play a role in the economic feasibility of succinate production. Abundance, cost, and availability of substrates are some of the important factors that make a substrate more or less attractive for production. Recently, the focus has been more on using renewable sources as substrates to lower the costs of production. Succinate production has been studied from varied substrates such as glucose, sucrose, lactose, fructose, glycerol, lignocellulosic hydrolysates, whey, cassava, and sugarcane molasses (7-15). 1.3 Applications for succinate and its derivatives Applications for succinate and its derivatives are many. Succinate is used to make a wide variety of products such as paints, coatings, adhesives, sealants, foods and flavors, cosmetics, nylons, industrial lubricants, phthalate-free plasticizers, dyes and pigments, as well as pharmaceuticals compounds (16). As bio-based succinate becomes more inexpensive, it can be used as a building block for a variety of chemicals such as adipic acid, 1,4-butanediol, γ-butyrolactone, tetrahydrofuran, n-methylpyrrolidone, poly-butylene succinate, 2-pyrrolidone, and polyamides (17, 18). About two-thirds of the bio-based succinate produced by year 2020 is expected to be used as an intermediate for producing 1,4-butanediol, tetrahydrofuran, and polyesters (19). 4 1.4 Succinate-producing microorganisms Succinate is an important metabolite in cellular metabolism, as an intermediate metabolite in the tricarboxylic acid (TCA) cycle and the glyoxylate shunt. Some microbes also produce it as an end product during anaerobic fermentations. Most bacterial species that naturally produce large amounts of succinate, such as Actinobacillus succinogenes, Mannheimia succiniciproducens, and B. succiniciproducens were isolated from the cow rumen (20-22). These natural succinate producers have been studied and engineered to further enhance succinate production. E. coli has been engineered for succinate production by mimicking the metabolism of these natural producers. E. coli, Corynebacterium glutamicum, and S. cerevisiae have also been engineered for succinate production by disrupting the TCA cycle and forcing flux through the glyoxylate shunt. Huge advances have been made in both natural producers and engineered strains towards succinate production. 1.4.1 Native succinate producers 1.4.1.1 Anaerobiospirillum succiniciproducens Anaerobiospirillum succiniciproducens was one of the first bacterial species studied for succinate production. When grown at pH 6.2 and high CO2 concentrations, A. succiniproducens produced high amounts of succinate, as opposed to lactate (23). Continuous cultures of A. succiniciproducens grown on glycerol (supplemented with yeast extract), yielded 1.17 mol succinate per mol glycerol with a productivity of 2.1 g L-1 h-1. Using a three-stage continuous cell recycle bioreactor, the succinate yield increased to 1.35 mol mol-1 glucose. Production rate and titer were 10.4 g L-1 h-1 and 83 g L-1, respectively (24). 5 Several features make this Gram-negative species unfit for industrial succinate production, though. It is highly sensitive to oxygen, making it difficult to handle. It is known to cause rare, but potentially lethal cases of bacteremia and diarrhea in humans (25), and no genetic tools are available, making it an unsuitable host for engineering. 1.4.1.2 Actinobacillus succinogenes A. succinogenes is a gram-negative, non-motile, osmotolerant, capnophilic, facultatively anaerobic bacterium that was isolated from a bovine rumen. It has the ability to grow on a broad range of substrates, such as glucose, lactose, xylose, arabinose, fructose, and glycerol (20). It belongs to the Pasteurellaceae family, along with Haemophilus, Aggregatibacter, Mannheimia, and Pasteurella. A. succinogenes’s fermentation pathways are well studied. A. succinogenes produces phosphoenolpyruvate (PEP) from glucose via the Embden-Meyerhof Parnas pathway. PEP is then either dephosphorylated to pyruvate by pyruvate kinase, or converted into oxaloacetate by PEP carboxykinase (PEPCK, encoded by pckA) (26). These two branches are interconnected by two decarboxylating enzymes, malic enzyme and oxaloacetate decarboxylase (27). Oxaloacetate is further reduced to malate, malate to fumarate, and fumarate is eventually reduced to succinate (Figure 1.2). A. succinogenes has an incomplete TCA cycle that lacks citrate synthase and isocitrate dehydrogenase (28). It also lacks the glyoxylate shunt (27). Metabolic flux analyses and genome annotation have led to a deeper understanding of A. succinogenes’s pathways and metabolism (27, 29, 30). A. succinogenes is capable of natural transformation and a knockout method has been developed (31). A. succinogenes does not ferment glycerol, but is can grow on glycerol by respiration. With nitrate as the terminal electron acceptor A. succinogenes produced acetate and CO2 as the main 6 products, together with small amounts of succinate, but no ethanol. In contrast, succinate became the main secreted product of glycerol-grown cultures with dimethyl sulfoxide (DMSO) as the electron acceptor, with a yield of 59% of the maximum theoretical yield, with formate, acetate, and CO2 as by-products. Similarly, succinate yields reached as high as 67% of the maximum theoretical yield under microaerobic conditions (1% O2, batch), with acetate, CO2 and formate as by-products. Deleting the pyruvate-formate lyase (ΔpflB) increased the succinate yield to 76% in microaerobic cultures (1% O2, batch) (15). Carvalho et al. grew A. succinogenes on glycerol with DMSO as the external electron acceptor under fed-batch conditions, and obtained a succinate titer of 49.6 g L-1 and a succinate yield of 0.5 mol mol-1 glycerol (32). The Nicol group developed a novel extended recycle biofilm reactor (continuous system) for fermentative succinate production. Using glucose as the substrate and CO2 to maintain anaerobic conditions, fermentations were carried out with media supplemented with yeast extract and corn steep liquor. At a 0.56 h-1 dilution rate, the succinate yield reached 1.05 mol mol-1 glucose with a 6.35 g L-1 h-1 productivity (33). The succinate yield was shown to be an increasing function of glucose consumption in that system, and the succinate yield could be increased to 1.39 mol mol-1 glucose with a 48.5 g L-1 succinate titer (34). In a separate study, Guettler et al. overexpressed the native glucose-6-phosphate dehydrogenase (G6PDH) in a pyruvate–formate lyase mutant strain to direct more flux through the pentose phosphate pathway. Under anaerobic conditions, the recombinant strain produced succinate with productivity, titer, and yield of 2.19 g L-1 h-1, 100.6 g L-1, and 1.36 mol mol-1 glucose, respectively (35). Vlysidis et al. first conducted batch experiments using a wide range of initial glycerol concentrations to test substrate inhibition on the process. One of the best results was found with a starting concentration of 37 g L-1 glycerol and yeast extract supplementation. These batch 7 cultures produced a succinate titer of 29.3 g L-1, a productivity of 0.27 g L-1 h-1, and a yield of 0.62 mol mol-1 glycerol. Next Vlysidis et al. developed an unstructured model by fitting a set of kinetic equations to the experimental data they found in the batch experiments done in small anaerobic reactors. This model was developed to predict experimental behavior and was validated by performing experiments in a scaled up bench top reactor. The model was able to predict the bench-top experiments without any additional fitting for a variety of conditions and can safely be used for future experiments (36, 37). An unstructured model was also developed to predict A. succinogenes fermentations on mixtures of C5 and C6 sugars found in spent sulfite liquor (38). Batch anaerobic fermentations were carried out using a mixture of xylose (72.6%), galactose, glucose, mannose, and arabinose with yeast extract. Succinate was produced with a titer of 27.4 g L-1, productivity of 0.45 g L-1 h-1, and yield of 0.7 g g-1 total sugars. Simulations were carried out using the kinetic parameters obtained from these experimental studies. Scaled up experiments using a 2-L lab-scale bioreactor using varying mixed sugar concentrations (15-50 g L-1) were performed to validate the predictive nature of their model. The R2 value of the model validations with the experimental results was 0.93, indicating that this model can effectively predict batch fermentations (38). The Bechkam group has studied the behavior of A. succcinogenes on xylose-enriched hydrolysates, a feedstock being widely studied as a substrate for production of value-added chemicals. This group grew A. succinogenes on deacetylated dilute acid-pretreated corn stover hydrolysate and obtained a succinate yield of 0.74 g g-1 sugars. The maximum succinate productivity and titer were 1.27 g L-1 h-1 and 42.8 g L-1, respectively (39). More recently, they have made ΔackA, ΔpflB and ΔpflBΔackA strains of A. succinogenes. When grown on mock biomass hydrolysates all strains showed a decrease in succinate titer, yield and productivity 8 compared to the wild-type strain indicating that, removal of heterofermentative pathways may not lead to increase in flux towards succinate (40). In the same study, they also overexpressed the PEP carboxykinase, malate dehydrogenase and fumarase in the wild-type strain as well as in all three of the above strains. They did not see any significant increase in yield, productivity and titers for either of ΔackA, ΔpflB and ΔpflBΔackA strains compared to the wild-type. All three of the wild-type strains expressing PEP carboxykinase, malate dehydrogenase and fumarase did show a slight increase in succinate yield and titer as compared to the wild-type strain. It is interesting to note that acetate accumulation was seen after a lag period in all ackA mutant backgrounds, suggesting an alternate route to acetate in A. succinogenes. Many other studies have tested for succinate production by A. succinogenes on different renewable carbon sources, such as straw hydrolysate, crop stalk waste, corn stover, cheese whey, and sugarcane bagasse (12, 41-44). 1.4.1.3 Mannheimia succiniciproducens In 2002, Lee et al. reported the isolation and characterization of a novel succinateproducing Pasteurellaceae species, Mannheimia succiniciproducens MBEL55E, from cow rumen. M. succiniciproducens produces succinate as its major fermentation product, along with acetate, formate, and lactate as by-products (22). M. succiniciproducens’s genome was sequenced and its metabolism was extensively studied (45). Genome-based metabolic engineering of M. succiniciproducens was carried out to study its metabolism and increase succinate production (45). In-silico flux analysis identified PEPCK as the major CO2-fixing enzyme instead of PEP carboxylase. PEPCK is the anaplerotic enzyme in the reductive TCA, 9 thus playing an important role in succinate production. Other enzymes responsible for formation of by-products during fermentation were also identified. During anaerobic fermentation on glucose in batch conditions, the wild-type strain produced 10.5 g L-1 of succinate, 4.9 g L-1 of acetate, 4.1 g L-1 of formate, and 3.5 g L-1 of lactate (45). LPK7 is a metabolically engineered strain of M. succiniciproducens, carrying ldhA, pflB, pta, and ackA deletions (45). In fed-batch cultures this strain produced succinate with yield, productivity, and titer of 1.16 mol mol-1glucose, 1.8 g L-1 h-1, and 52.4 g L-1, respectively. With glycerol as the carbon source, the PALK strain (ΔldhA Δpta-ackA) had a succinate yield of 0.88 mol mol-1 glycerol with an overall productivity of 0.13 g L-1 h-1. The PALK strain grown in fedbatch conditions using glucose and glycerol as co-substrates along with 6.84 M magnesium hydroxide and 1.57 M ammonia for pH control gave a succinate yield of 1.15 mol mol-1 glucose equivalent, with an overall productivity of 3.5 g L-1 h-1 and a titer of 90.7 g L-1, together with acetate (2.3 g L-1) and pyruvate (4.0 g L-1) as byproducts (46). Recent simulations (47) based on omics studies and metabolism reconstructions predicted that sucrose and glycerol used as co-substrates would increase succinate production while reducing by-product formation. In this study, the PALKF (ΔldhA Δpta-ackA ΔfruA) strain was constructed to divert the majority of the flux towards succinate production, minimize by-product formation, and deregulate catabolite repression. The PALKF strain was grown in fed-batch conditions as low and medium density cultures. The succinate yields obtained were 1.56 and 1.64 mol mol-1 glucose equivalent with overall productivities of 2.50 and 6.02 g L-1 h-1 for low and medium density cultures, respectively. The PALKG strain (expressing the allosteric inhibition-free E. coli glpK22) (48, 49) was also developed, which gave a higher productivity (3.34 g L-1 h-1) than the PALKF strain, but produced more by-products. A membrane cell recycle 10 bioreactor system was also developed in which the PALKF strain had a succinate productivity of 38.6 g L-1 h-1 (47). Overall, Dr. Sang Yup Lee’s group was able to make a homo-succinate production possible with high productivities using sucrose and glycerol as co-substrates. 1.4.1.4 Basfia succiniciproducens B. succiniciproducens DD1 was isolated from the bovine rumen and characterized as another succinate-producing Pasteurellaceae species (50). B. succiniciproducens DD1 and M. succiniciproducens MBEL55E had similar genome sizes, and their homologous proteins shared 95% similarity (21). When grown on glucose or sucrose as the carbon source, B. succiniciproducens DD1 produced succinate with a 5.8 g L-1 titer, a 1.5 g L-1 productivity, and a 0.6 g g-1 sucrose yield. With glycerol as the carbon source the succinate titer was higher, at 8.4 g L-1, with a productivity of 0.9 g L-1 h-1 and a yield of 1.2 g g-1 glycerol (50). A continuous cultivation process was developed on crude glycerol supplemented with yeast extract and was maintained for 80 days. At a dilution rate of 0.018 h-1 the succinate titer was 5.21 g L-1, the productivity was 0.094 g L-1 h-1, and the yield was 1.02 g g-1 glycerol (50, 51). Metabolic flux analysis showed unwanted fluxes through pyruvate-formate lyase and lactate dehydrogenase. A ΔldhA ΔpflD strain was developed that showed a succinate yield (1.08 mol mol-1 glucose) reaching 62% of the maximum theoretical yield (52). Succinity has a patent for the engineered strain and is using it for commercial scale production of succinate in its facilities. 11 1.4.2 Engineered succinate producers 1.4.2.1 Escherichia coli E. coli is metabolically versatile and can consume glucose both anaerobically and aerobically. Under anaerobic conditions it carries out a mixed acid fermentation when grown on glucose to produce formate, lactate, ethanol, acetate, and succinate, where succinate is only a minor fermentation product (53). Under aerobic conditions it only produces succinate if the glyoxylate shunt is operational (54). The theoretical maximum yield for succinate in E. coli grown fermentatively on glucose is 1 mol mol-1 glucose (54). Wild-type E. coli produces only 0.11 mol succinate per mol glucose in these conditions (55). The succinate yield on glucose is generally limited by the availability of reducing equivalents (54). Although E.coli is not a natural succinate producer, strategies have been developed to improve succinate production in E. coli by engineering, evolution, or optimizing the production conditions. In one of the earlier studies by Millard et al., overexpressing PEP carboxylase (PEPC) increased succinate production 3.75-fold, but concomitantly decreased glucose uptake (56). A wild-type E. coli strain overexpressing a heterologous pyruvate carboxylase (from Rhizobium etli) produced 1.77 g L-1 succinate compared to 1.18 g L-1 by the wild-type strain (57). Strain AFP111 carrying knockout mutations in pflB and ldhA was unable to ferment glucose (53). A spontaneous mutation in the ptsG gene restored the ability of AFP111 derivative NZN111 to ferment glucose while producing 1 mol of succinate, as well as 0.5 mol of acetate and ethanol each per mol of glucose (58). Strain AFP111(pTrc99A-pyc), which overexpresses the R. etli pyruvate carboxylase, produced 99.2 g L-1 of succinate with yield and productivity of 1.74 mol mol-1 glucose and 1.3 g L-1 h-1, respectively, when grown in dual phase conditions (cultures grown aerobically to generate biomass, then transferred to anaerobic conditions conducive to succinate production) (59). Kim 12 et al., 2004 demonstrated that PEPCK can replace PEPC as the PEP-carboxylating enzyme in E. coli. Overexpressing the A. succinogenes 130Z PEPCK in E. coli K-12 ppc::kan increased succinate production 6.5-fold (60). Dr. Lonnie Ingram’s group combined metabolic evolution and gene knockouts to develop a good biocatalyst for succinate production. First the genes responsible for the formation of major by-products were knocked out, namely acetate kinase (ackA), lactate dehydrogenase (ldhA), and alcohol dehydrogenase (adhE). This strain, called KJ012, grew poorly on glucose minimal medium under anaerobic conditions. In this strain, however, ATP production was coupled to succinate production. Strain KJ012 was evolved to grow anaerobically in glucose minimal medium for over 2,000 generations to allow for growth-based selection. The evolved strain grew much better on glucose minimal medium under anaerobic conditions, and produced more succinate than its parent strain. Additionally, the genes focA and pflB were deleted as well, increasing the succinate production levels even higher. This new strain, KJ060, produced a succinate yield of 1.41 mol mol-1 glucose, with productivity and titer of 0.9 g L-1 h-1 and 86.6 g L-1 respectively. The succinate yield reached 1.61 mol mol-1 glucose when higher inoculums were used (61). KJ134 (ΔldhA ΔadhE ΔfocA-pflB ΔmgsA ΔpoxB ΔtdcDE ΔcitF ΔaspC ΔsfcA Δpta-ackA) is a derivative of KJ060. This strain had additional deletions in pta, tdcE, and tdcD to further reduce acetate production. The carbon flux to oxaloacetate was also increased by knocking out genes coding for aspartate aminotransferase (aspC) and NADP-linked malic enzyme (sfcA). KJ134 produced 1.53 mol succinate per mol of glucose with a productivity of 0.75 g L-1 h-1 and a titer of 71.56 g L-1 (62). Additionally, strain KJ134 showed an 80% reduction in acetate production as compared to KJ073 (ΔldhA, ΔadhE, ΔackA, ΔfocA, ΔpflB, ΔmgsA, 13 ΔpoxB). When grown on 2% glycerol, KJ073 produced 15 g L-1 succinate from 16.9 g L-1 glycerol, a yield of 89% of the theoretical maximum (61). During the evolution of strains KJ060 and KJ073, the authors noticed changes in metabolism that led to increased succinate production and ATP yield per glucose (63). They found a spontaneous mutation in the promoter of pckA, encoding gluconeogenic PEPCK, that allowed PEPCK to replace PEPC as the main carboxylating enzyme. In addition, glucose uptake through the PEP-dependent phosphotransferase system had been inactivated (spontaneous mutation) and replaced by galactose permease and glucokinase. These E. coli strains now had a pathway similar, for succinate production, to the natural succinate-producing organisms found in cow rumens. The Ingram group confirmed that these changes led to higher succinate amounts by engineering an E. coli strain with the mutations seen above. Strain XZ647, with only two mutations (ptsI truncation and constitutive pckA), was able to produce succinate with a yield of 0.89 mol mol-1 of glucose. An additional pflB deletion yielded strain XZ721, whose succinate yield increased to 1.25 mol mol-1glucose (64). When grown on glycerol, strain XZ721 had a yield of 0.8 mol mol-1 of glycerol (63, 65). Dr. Ramon Gonzalez’s group also engineered E. coli for succinate production from glycerol. Their approach included blocking the pathways to major by-products and overexpressing a heterologous pyruvate carboxylase to increase succinate production by driving the flux from pyruvate to oxaloacetate. The final strain, ΔadhE Δpta ΔpoxB ΔldhA Δppc [pZSpyc], overexpressed the Lactobacillus lactis pyruvate carboxylase and produced succinate with a yield of 0.54 mol mol-1 glycerol in microaerobic conditions (14). 14 1.4.2.2 Corynebacterium glutamicum C. glutamicum is a gram-positive, non-motile, spore-forming, facultatively anaerobic bacterium that belongs to the Actinomycetes subdivision of eubacteria. It has been widely studied and used as an industrial strain for the production of amino acids and other organic acids (66). Okino et al. observed that C. glutamicum incubated in glucose-mineral medium under oxygen-limited conditions (without growth) produced succinate, lactate, and acetate (67). Increased bicarbonate concentrations led to an increased succinate yield and a decreased lactate yield. The ability of this species to produce organic acids in arrested growth conditions allowed for a bioreactor design in which high cell density led to a high succinate volumetric productivity of 11.7 g L-1 h-1 (67). The C. glutamicum strain ΔldhA-pCRA717, a lactate dehydrogenase knockout that overexpresses pyruvate carboxylase, produced 146 g L-1 succinate in 46 h and the succinate and acetate yields were 1.4 and 0.29 mol mol-1 glucose, respectively (68). The succinate productivity was 3.2 g L-1 h-1. Litsanov et al. constructed the BOL-1 strain, in which all known pathways leading to acetate and lactate synthesis were deleted, and the BOL-2 strain where the pyruvate carboxylase gene was integrated chromosomally into the BOL-1 strain. Integrating the Mycobacterium vaccae NAD+-coupled formate dehydrogenase into the BOL-2 chromosome gave strain BOL-3. Additionally, a metabolic blockage of glycolysis caused by NADH inhibition of GAPDH activity in glucose-grown BOL-3 was relieved by overexpressing the native glyceraldehyde-3-phosphate dehydrogenase gene on a plasmid. This final strain grown in the presence of glucose and formate in fed-batch conditions produced 134 g L-1 of succinate in 53 h. The succinate yield was 1.67 mol mol-1 glucose with very little accumulation of other by-products (0.1 mol mol-1 glucose) (69). 15 Figure 1.2 Metabolic map of Actinobacillus succinogenes. AcCoA, acetyl CoA; CoA, Coenzyme A; OAA, oxaloacetate; and PEP, phosphoenolpyruvate. Another study came up with a dual synthesis approach for succinate production without using formate. The glyoxylate pathway citrate synthase genes were overexpressed to direct more carbon toward the glyoxylate pathway. The succinate exporter SucE was overexpressed as well. This strain produced 109 g L-1 succinate with an overall volumetric productivity of 1.11 g L-1 h-1 and a yield of 1.32 mol mol-1 glucose (70). All these studies were carried out under anaerobic conditions. More recently, a few studies have focused on aerobic production of succinate in C. glutamicum. Strain BL-1 (pAN6-pycP458Sppc) has in-frame deletions of pqo, pta-ackA, sdhCAB, and cat, and overexpresses the pyruvate carboxylase (pyc) and PEPC (ppc) genes. This strain is the first known to produce succinate aerobically in minimal glucose medium. Grown under 16 conditions that decoupled succinate production from growth, this strain produced a titer of 10.6 g L-1 succinate with a yield of 0.45 mol mol-1 glucose (71). Strain BL-1(pVWEx1-glpFKD) overexpresses the E. coli glycerol facilitator, glycerol kinase, and glycerol-3-phosphate dehydrogenase. When grown aerobically on glycerol minimal medium, it produced 9.3 g L-1 of succinate with a volumetric productivity of 0.42 g L-1 h-1 (72). In another study, B. subtilis acetyl-coA synthetase was overexpressed on a plasmid in the base strain ZX1 to recycle the carbon in acetate. The ZX1 strain has deletions of ldhA, pqo, cat, and pta genes along with replacement of the native promoters of the pyc and ppc with the sod promoter. Strain ZX1(pEacsA) did not secrete any acetate, and the succinate yield was 0.5 mol mol-1 glucose. Overexpressing citrate synthase (encoded by gltA) led to an additional 22% increase in succinate yield. When strain ZX1(pEacsAgltA) was grown in fed-batch conditions on glucose it produced 28.4 g L-1 of succinate with a volumetric productivity of 0.41 g L-1 h-1 and a yield of 0.63 mol mol-1 glucose. This study, however, did use rich medium components in their production medium (73). A recent study reported the development of an engineered strain of C. glutamicum, S071 (ΔldhA Δpta-ackA ΔactA ΔpoxB pycP458S Ptuf::ppc Δpck_Ptuf::pckG ΔptsG) able to produce 152.2 g L-1 of succinate with a yield of 1.67 mol mol-1 glucose by over-expressing the transcriptional regulator NCgl0275. Over-expressing NCgl0275 allowed the release of an end-product inhibition due to succinate by increasing the glucose consumption rate (74). 1.4.2.3 Saccharomyces cerevisiae S. cerevisiae is the only yeast that has been well studied for succinate production. Succinate is the main component that imparts flavor to Sake during fermentation (75). Arikawa et al. 17 studied the metabolic pathways leading to succinate production in S. cerevisiae and established that succinate could be produced by either n-ketoglutarate oxidation or fumarate reduction. In one of the first engineering studies of S. cerevisiae, inactivation of succinate dehydrogenase (sdh and sdh1b deletions) generated a strain with reduced malate productivity, and a succinate productivity about double that of the wild-type strain (76). Another group deleted sdh1, sdh2, idh1, and idp1 to completely abolish isocitrate dehydrogenase and succinate dehydrogenase activities, diverting the TCA cycle flux into the glyoxylate pathway, and increasing succinate production. In glucose-shake flask cultures, this strain produced 3.62 g L-1 of succinate with a yield of 0.11 mol mol-1 glucose (77). Ito et al. (78) deleted sdh1 and sdh2 (encoding succinate dehydrogenase subunits) to allow for aerobic succinate production. To further increase succinate production they deleted the ethanol-producing genes, adh1, adh2, adh3, adh4, and adh5 (encoding alcohol dehydrogenases). This strain, S149sdh12, produced 20-fold more succinate than the control strain, with a yield of 0.22 mol mol-1 glucose. Since succinate was accumulating intracellularly, they overexpressed the Schizosaccharomyces pombe malic acid transporter (mae1) in this engineered strain and further increased the succinate yield to 0.24 mol mol-1 glucose. Yan et al. engineered S. cerevisiae for succinate production through the reductive branch of the TCA cycle. In S. cerevisiae, fum1 encoding fumarate hydratase irreversibly converts fumarate to malate, which poses a problem for reductive succinate production. The authors deleted fum1 and overexpressed E. coli fumC in its place, in a pyruvate decarboxylase-deficient (TAM) strain. The pyc2, mdh3, and frd1 genes were also overexpressed to increase pyruvate decarboxylase, malate dehydrogenase, and fumarate reductase activity, respectively. GPD1 (glycerol-3-phosphate dehydrogenase) is responsible for NADH-dependent glycerol production 18 in aerobic conditions (79). Since production of 1 mol of succinate consumes 2 mol of NADH through the reductive TCA branch, the authors suspected that inactivating GPD1 would potentially increase succinate production. When grown in a bioreactor with optimal CO2 (10%) and medium conditions, the final strain, PMCFfg, produced succinate with a titer of 13 g L-1 and a yield of 0.21 mol mol-1 glucose at a pH of 3.8 (79). Figure 1.3 Pathways to succinate production by mixed acid metabolism in bacteria. Abbreviations are the same as in Figure 1.2. 19 Table 1.1 Succinate-producing strains with their yields, productivities, and titers Strain Conditions/Substrate Succinate Yield (mol mol-1) Productivity (g L-1 h-1) Other products Reference 31.7 24.4 49.6 48.5 0.75 0.74 0.50 1.39 0.14 2.13 0.96 ND A, F A, F A, F A, F (15) (32) (32) (34) 90.7 78.4 1.15a 1.64a 3.49 6.03 P, A P, A (46) (47) 13.2 1.22a 38.6 P, A (47) 69.8 1.70 2.91 46.3 1.33 ND A, P, L, M, F A (78, 80, 81) (80) 99.2 86.6 11.7 71.5 8.75 1.74 1.41 0.89 1.53 0.77 1.3 0.9 ND 0.75 ND A, E A, M, L ND A, P, M ND (59) (61) (61) (62) (65) 146 134 1.4 1.67 3.2 2.48 A, L, M, P K, M, A, F, P (68) (69) ND 3.62 13.0 0.24 0.11 0.21 ND ND ND G, E, Fum K, P, C G, P (78) (77) (79) Titer (g L-1) Actinobacillus succinogenes 130Z 130Z 130Z 130Z D, CC, glycerol, Mi D, glycerol, B, DMSO, An, YE D, glycerol, FB, DMSO, An, YE D, glucose, YE, CSL, An LU15224 Mannheimia succiniciproducens D, FB, An, glucose + glycerol D, FB, An, sucrose + glycerol, high cell concentration (Initial OD 600 of 9.03) D, MRCB (dilution rate of 2.93l h-1), An, sucrose + glycerol Basfia succiniciproducens DD1 D, B, An, glycerol + maltose LU15224 pJFF224 (icl ms Y.m.) D, B, An, glucose, YE PALK PALKF PALKF Escherichia coli AFP111(pTrc99A-pyc) KJ060 KJ134 XZ465 D, FBc, YE, T D, B, glucose D, B, glycerol D, B, glucose D, B, glycerol Corynebacterium glutamicum ΔldhA-pCRA717 BOL-3/pAN6-gap (Δcat, Δpqo, Δpta-ack, Δldh, pycP4585, fdh) Db, B, OD, glucose Db, FB, An, glucose + formate S149sdh12/pNV11-mae1 Δsdh1Δsdh2Δidh1Δidp1 PMCFfg D, B, Ae, glucose D, B, Ae, glucose D, B, Ae, glucose Saccharomyces cerevisiae Abbreviations: A, acetate; Ae, aerobic; An, anaerobic; B, batch; C, Citrate; CC, continuous culture; CSL, corn steep liquor; D, defined; DMSO, dimethyl sulfoxide; E: ethanol; F, formate; Fum, fumarate; FB, fed-batch; G, Glycerol; K, α-ketoglutarate; L, lactate; M, Malate; Mi, microaerobic; MCRB, membrane cell recycling bioreactor; ND: Not determined; OD, oxygen deprivation (dissolved oxygen lower than 0.01 ppm); P, pyruvate; T, tryptone; VMD, vacuum membrane distillation; and YE, yeast extract. a mol mol-1 glucose equivalent b aerobic cell propagation and transfer to anaerobic succinate production conditions c dual phase fed-batch where cells were grown in aerobic conditions, then oxygen-free CO2 was sparged at transition point 20 1.5 C4-dicarboxylate transporters Engineering of microbial species for succinate production has mostly focused on targets within intracellular metabolic pathways. Very few engineering studies have focused on expressing succinate transporters to increase succinate production. E. coli and C. glutamicum are two commercially relevant strains in which succinate transporters have been characterized and used to increase succinate production. C4-dicarboxylates such as succinate, fumarate, and malate are intermediates in the TCA cycle and serve as carbon/energy sources during aerobic growth and therefore require uptake systems (82). During anaerobic growth, all C4-dicarboxylates other than succinate are used as electron acceptors in fumarate respiration. In contrast, succinate is exported via antiport against the uptake of malate or fumarate (82). Many different types of transporters are involved in C4-dicarboxylate transport and, with a few exceptions, most studies to identify these transporters have focused on E. coli and on C4-dicarboxylate uptake. 1.5.1 E. coli C4-dicarboxylate transporters So far, E. coli has seven known C4-dicarboxylate transporters —DauA, DctA, DcuA, DcuB, DcuC, YjjPB, and CitT. Unidirectional uptake of fumarate, succinate, malate, and aspartate takes place under aerobic conditions, while exchange, uptake, and efflux of these C4dicarboxylates take place in anaerobic conditions (83). DauA and DctA are primarily known to function in aerobic conditions while all others are active in anaerobic conditions. DauA (dicarboxylic acid uptake system A) has been characterized as a succinate uptake transporter active under acidic pH conditions (84). DctA, a dicarboxylate amino acid-cation symporter family (DAACS) transporter is used for aerobic growth in many bacteria such as E. coli, Bacillus subtilis, Rhizobium leguminosarum, and C. glutamicum (85-87). It mediates the uptake of 21 succinate, malate and fumarate in E. coli under aerobic conditions. DcuA, DcuB, and DcuC are antiporters found only in anaerobic and facultatively anaerobic bacteria and are involved in fumarate respiration. The DcuAB and DcuC families are independent from each other, but members of both families are capable of uptake, exchange, and efflux of C4-dicarboxylates. The main functions of DcuA, DcuB, and DcuC are still not completely elucidated. An early study determined that DcuAB carriers mostly functioned as uptake or exchange carriers whereas DcuC carriers likely mostly operated as efflux carriers (82). Still, an E. coli DcuC mutant grown anaerobically was not affected in succinate production, indicating that DcuC is not the main succinate exporter in anaerobic conditions (88). Another E. coli study showed that individual deletions of dcuB and dcuC decreased succinate titer by 15 and 11%, and a ΔdcuBΔdcuC double mutant showed a 90% decrease in succinate titer in a glucose fermentation. Modulating the expression of these two transporters in E. coli also increased the succinate titer by 34% (89). Another family of transporters involved in C4-dicarboxylate transport is the TRAP (tripartite-ATP independent periplasmic) family. In TRAP carriers, the driving force for solute accumulation is an electrochemical ion gradient (90). TRAP carriers consist of two membrane integral proteins (DctQM) and a periplasmic solute-binding protein (DctP). The first characterized TRAP transporter was the R. capsulatus DctPQM system and homologs of this system were found to be present in a wide range of bacteria and archaea (90). The TRAP family carriers are involved in solute uptake and are considered unidirectional (91). For this reason, it is unlikely that they could be involved in succinate efflux. Another known E. coli C4-dicarboxylate transporter is CitT, however CitT is not restricted to only C4-dicarboxylate transport. E. coli CitT is involved in citrate uptake in exchange for succinate, fumarate, or tartarate under anaerobic conditions (82, 92). More recently, a study by 22 Fukui et al. (88) demonstrated that E. coli YjjP and YjjB form a single complex involved in succinate export under both aerobic and anaerobic conditions. A YjjPB mutant showed a 70% decrease in succinate production suggesting that YjjPB is involved in succinate export. YjjPB were found to contain different structures and conserved domains when compared to other known C4-dicarboxylate transporters (88). 1.5.2 C4-dicarboxylate transporters in other succinate producers BlastP searches with E. coli YjjP in A. succinogenes, M. succiniciproducens, and E. aerogenes predicted homologs with 50% (87%), 54% (89%) and 88% (90%) identity (similarity), respectively (88). Similar BlastP searches with E. coli YjjB identified homologs in A. succinogenes, M. succiniciproducens, and E. aerogenes with 50% (81%), 52% (83%), and 83% (97%) identity (similarity) (88). Whether these homologs function in succinate export is unknown yet. The C. glutamicum proteome does not contain any homologs of the E. coli Dcu family succinate transporters. In a study of C. glutamicum transmembrane proteins, 57 membrane proteins with at least four trans-membrane domains were found to be significantly similar to A. succinogenes and M. succiniciproducens proteins. Eliminating those with predicted or known functions brought the list of candidates down to twenty. Construction of deletion mutants of twelve of these candidates allowed the identification of SucE. The sucE mutant produced 70% of the wild-type levels of succinate and thus may not be the main succinate exporter in anaerobic conditions. SucE is a secondary carrier like the Dcu family carriers in E. coli (93). Another study identified SucE through transcriptomics studies for C. glutamicum grown in anaerobic and microaerobic conditions (94). Unlike the Dcu family carriers SucE is only involved in succinate 23 export (93). In another study, SucE-overexpressing C. glutamicum produced 109 g L-1 of succinate and a yield of 1.32 mol mol-1 glucose as mentioned above (70). SucE homologs were found in E. coli, A. succinogenes, and M. succiniciproducens (93), but their function has not been tested. The only identified succinate transporter in A. succinogenes is Asuc_0304. It was identified by complementation studies done in a C4-dicarboxylate transport deficient mutant E. coli strain IMW213. Asuc_0304 was identified to be an uptake C4-dicarboxylate transporter capable of fumarate or malate uptake with Na+ symport, although it may not be the sole or main transporter under aerobic conditions (95). 1.6 sRNAs as synthetic regulators for metabolic engineering Industrially-relevant microorganisms have been subject to heavy modifications to optimize yield, productivity, and titers of their products. Most engineering strategies are aimed at identifying target genes and either knocking them out or over-expressing them. Many groups also take a rational approach for evolving strains for growth under certain conditions. However, it is often not easy to manipulate microorganisms and therefore conduct any type of metabolic engineering. In addition, these methods can only target a limited number of genes at a time and are therefore unsuitable for applications on a genomic scale (96). Another disadvantage of overexpressing genes is the limited control over expression level and/or stability in the particular organism. There is a growing need for strategies that enable researchers to modulate gene expression or repression at the translational level. Synthetic sRNAs are gaining tremendous momentum in research, as using them would allow us to overcome many of the challenges listed above. One of the major advantages of 24 synthetic sRNAs is the possibility of designing them with a rational approach, making it unnecessary to build individual strains to test the intended phenotypes of the knockout or overexpression of genes. A recent explosion of studies has started to decipher the different ways in which sRNAs modulate different genes in nature. The increased understanding of how sRNAs function has led to an interest in using them as synthetic regulators by mimicking in vivo gene regulation to target genes of interest. Synthetic small RNAs have been designed for E. coli in several studies (42, 96-102). Desai et al. (103) demonstrated that expressing non-coding antisense sRNAs (asRNAs) complementary to the butyrate kinase, phosphotransbutyrylase, and phosphotransacetylase mRNAs inhibited these enzymes’ activities and increased acetone and butanol production. Kim et al. designed antisense RNAs to lower the expression of acetate kinase and phosphotransacetylase. This strategy partially reduced the mRNA levels of the target genes and blocked the synthesis of the enzymes (97). Nakashima et al. designed an asRNA with pairedend termini for conditional silencing of acetate kinase and phosphotransacetylase. The pairedend termini increased the asRNA stability and improved the conditional silencing in E. coli (98, 99). Negrete et al. also demonstrated that overexpressing the GadY sRNA in E. coli decreased acetate production, which in turn decreased the inhibitory effect acetate had on the strain’s growth (99). Several of the early discovered sRNAs were found to be Hfq-dependent. Hfq is an RNAbinding protein that acts as a chaperone by enabling the binding of sRNAs to their target RNAs. Hfq is also known to stabilize sRNAs and their target mRNAs, thereby increasing their half-lives (104). E. coli Hfq-dependent sRNAs have an mRNA base-pairing region, Hfq-binding region, and a Rho-independent terminator (Figure 1.4). The Hfq-binding region contains a consensus secondary structure that provides a scaffold for Hfq to bind. In one of the first studies that 25 implemented sRNAs for gene repression, artificial sequences complementary to the ribosomal binding sites and start codons of ipp, ompA, and ompC were inserted in between the stem loops at the 5’ and 3’ termini of the natural hfq-dependent MicF sRNA. These synthetic sRNAs could inhibit the production of OmpC, OmpA, and Ipp (100). In another E. coli study the known sRNA RhyB (involved in regulation of iron metabolism) was overexpressed with an arabinose inducible promoter. Overexpressing RhyB during growth on glucose led to a 7-fold increase in succinate production without citrate production. RhyB was shown to regulate sdhCDAB and acnB, which encode succinate dehydrogenase and aconitase, respectively, in the TCA cycle (101). Overexpressing RhyB in the E. coli DALRA strain (engineered for 5-aminolevulinic acid production) allowed for 16% higher 5-aminolevulinic acid production than the parent strain (42). Na et al. designed 130 synthetic sRNAs by using two separate parts—the scaffold sequence and the mRNA or target-binding sequence. Using this approach the authors were able to substantially increase the production of cadaverine and tyrosine in E. coli. At high cell densities, the best engineered strain produced a tyrosine titer of 21.9 g L-1, whereas cadaverine production went up to 12.6 g L-1, compared to 9.6 g L-1 for the starting strain (96). 26 Figure 1.4 Hfq-dependent sRNA organization and structure in E. coli. Designing sRNAs in E. coli is much more straightforward now with the vast amount of knowledge gained over the past decade. However, for other organisms where sRNAs are still being identified and the Hfq-binding scaffold sequence is not clearly understood or identified, designing synthetic, Hfq-dependent sRNAs is much more difficult. To engineer B. subtilis, Liu et al. (105) overcame these hurdles by designing synthetic sRNAs based on the E. coli MicC scaffold, and by coexpressing them with E. coli Hfq in B. subtilis. With this approach, the authors repressed the expression of pfk (encoding 6-phosphofructokinase) and glmM (encoding phosphoglucosamine mutase) to increase N-acetylglucosamine production. Cho et al. used the same approach in Clostridium acetobutylicum PJC4BK to repress phosphotransacetylase expression. Strain PJC4BK (pPta-HfqEco) produced 16.9 g L-1 of butanol, which was higher than the starting strain as acetate production was inhibited (106). Research is underway to discover new sRNAs in many bacterial species. 27 1.7 Identification of small RNAs in Pasteurelleaceae Pasteurellaceae are a family of Gram-negative bacteria that are mostly commensals on the mucosal surfaces of birds and mammals. The Pasteurellaceae family consists of eighteen genera, some of the major ones being Haemophilus, Actinobacillus, and Pasteurella. Only a handful of papers mention sRNAs in Pasteurelleaceae. Subashchandrabose et al. demonstrated the role of Hfq in pathogenesis of Actinobacillus pleuropneumoniae. A Δhfq mutant of A. pleuropneumoniae was defective in biofilm formation and sensitive to oxidative stress; however not much more is known regarding the sRNAs involved (107). Three sRNAs have been identified in Aggregatibacter actinomycetemcomitans that are iron- and fur-regulated (108). Baddal et al. (109) identified seventeen novel putative sRNAs in intergenic regions, six of them homologous to known sRNA families in Haemophilus influenzae. HrrF, a fur-regulated sRNA was the first sRNA to be identified in any Haemophilus species (110). More recently, the Bazzolli group conducted an extensive study to identify small RNAs in A. pleuropneumoniae, an organism known to cause porcine pleuropneumonia. The authors used four different algorithms to identify small RNAs and were able to experimentally confirm seventeen of the 23 sRNAs found by all four algorithms by Northern blotting, RT-PCR, and RNA sequencing. They also found that the sequences of these seventeen sRNAs were well conserved in the species that are evolutionarily close to A. pleuropneumoniae (111). These early studies have shed some much needed light in this unexplored area of sRNAs in Pasteurellaceae. However, much remains to be discovered. 1.8 Introduction to chapters The main aims of this dissertation are to: 28 (i) provide insight into small RNAs and make progress in using small RNAs as a valuable tool for increasing succinate production in A. succinogenes. (ii) make a promoter library for A. succinogenes for tunability of gene expression (iii) carry out a proteomics- and transcriptomics-based search for succinate export transporters and see the effect of their overexpression on succinate production. Chapter 2 describes a proteomics- and transcriptomics-based search for succinate exporters in A. succinogenes. Transporters with the most hits in our omics data were further tested by overexpression for succinate production. Tunability for expressing these transporters was required and so far no such system was available in A. succinogenes. For this purpose, a reporter assay system was developed in A. succinogenes to test promoter strength. Several promoters were tested for strength to generate a library of promoters of increasing strengths. The four candidate succinate exporters identified were expressed under control of a few of the weaker promoters to identify the transporter expression levels that would not interfere with A. succinogenes growth and test their effect on succinate production. . Chapter 3 describes a sequencing study of A. succinogenes small RNAs conducted on cultures grown anaerobically on glucose and microaerobically on glycerol. The main aim of this study was to find Hfq-dependent small RNAs and compare their sequences to be able to design a synthetic small RNA for repression of mRNA in A. succinogenes. Chapter 4 describes a method for making markerless knockout mutants in A. succinogenes using natural transformation. Chapter 5 describes the work I did as a part of a previously published paper. I developed a continuous culture system for A. succinogenes growth on glycerol under microaerobic conditions. 29 In chapter 6, I summarize what was learned about succinate transporters and sRNAs in A. succinogenes. I also discuss future work for gaining more insight and further engineering of A. succinogenes for succinate production. 30 REFERENCES 31 REFERENCES 1. Werpy TA, Petersen G. 2004. Top value added chemicals from biomass for the U.S. Department of Energy (DOE) by National Renewable Energy Laboratory (NREL). 2. Patel MK, Crank M, Dornburg V, Hermann B, Roes L, Hüsing B, Overbeek L, Terragni FR, E. 2006. The BREW project: Medium and Long-term Opportunities and Risks of the Biotechnological Production of Bulk Chemicals from Renewable Resources– the Potential of White Biotechnology. Technology and Society (STS)/Copernicus Institute, Utrecht, the Netherlands. 3. Bozell JJ, Petersen GR. 2010. Technology development for the production of biobased products from biorefinery carbohydrates—the US Department of Energy’s “top 10” revisited. Green Chem 12:539-554. 4. WP 8.1. Determination of market potential for selected platform chemicals. Itaconic acid, succinic acid, 2,5-furandicarboxylic acid. Weastra, S. R. O. 5. Biddy MJ, Scarlata C, Kinchin C. 2016. Chemicals from Biomass: A Market Assessment of Bioproducts with Near-Term Potential. NREL (National Renewable Energy Laboratory, Golden, CO, USA. 6. www.myriant.com. Accessed 7. Sawisit A, Jantama SS, Kanchanatawee S, Jantama K. 2014. Efficient utilization of cassava pulp for succinate production by metabolically engineered Escherichia coli KJ122. Bioproc Biosyst Eng 38:175-187. 8. Chen C, Ding S, Wang D, Li Z, Ye Q. 2014. Simultaneous saccharification and fermentation of cassava to succinic acid by Escherichia coli NZN111. Bioresour Technol 163:100-105. 9. Chatterjee R, Millard CS, Champion K, Clark DP, Donnelly MI. 2001. Mutation of the ptsG gene results in increased production of succinate in fermentation of glucose by Escherichia coli. Appl Environ Microbiol 67:148-154. 10. Wang J, Zhu J, Bennett GN, San KY. 2011. Succinate production from sucrose by metabolic engineered Escherichia coli strains under aerobic conditions. Biotechnol Prog 27:1242-1247. 11. Liu R, Liang L, Cao W, Wu M, Chen K, Ma J, Jiang M, Wei P, Ouyang P. 2013. Succinate production by metabolically engineered Escherichia coli using sugarcane bagasse hydrolysate as the carbon source. Bioresour Technol 135:574-577. 32 12. Borges ER, Pereira Jr N. 2011. Succinic acid production from sugarcane bagasse hemicellulose hydrolysate by Actinobacillus succinogenes. J Ind Microbiol Biotechnol 38:1001-1011. 13. Samuelov NS, Datta R, Jain MK, Zeikus JG. 1999. Whey fermentation by Anaerobiospirillum succiniciproducens for production of a succinate-based animal feed additive. Appl Environ Microbiol 65:2260-2263. 14. Blankschien MD, Clomburg JM, Gonzalez R. 2010. Metabolic engineering of Escherichia coli for the production of succinate from glycerol. Metab Eng 12:409-419. 15. Schindler BD, Joshi RJ, Vieille C. 2014. Respiratory glycerol metabolism of Actinobacillus succinogenes 130Z for succinate production. J Ind Microbiol Biotechnol 41:1339-1352. 16. www.bio-amber.com. 17. Chen Y, Nielsen J. 2016. Biobased organic acids production by metabolically engineered microorganisms. Current Opin Biotechnol 37:165-172. 18. Ahn J-H, Sang B-I, Um Y. 2011. Butanol production from thin stillage using Clostridium pasteurianum. Bioresource Technol 102:4934-4937. 19. Cavani F, Albonetti S, Basile F, Gandini AA. 2016. Chemicals and Fuels from BioBased Building Blocks. Wiley Online Library. 20. Guettler MV, Rumler D, Jain MK. 1999. Actinobacillus succinogenes sp. nov., a novel succinic-acid-producing strain from the bovine rumen. Int J Syst Bacteriol 49:207–216. 21. Kuhnert P, Scholten E, Haefner S, Mayor D, Frey J. 2010. Basfia succiniciproducens gen. nov., sp. nov., a new member of the family Pasteurellaceae isolated from bovine rumen. Int J Syst Evol Microbiol 60:44-50. 22. Lee PC, Lee SY, Hong SH, Chang HN. 2002. Isolation and characterization of a new succinic acid-producing bacterium, Mannheimia succiniciproducens MBEL55E, from bovine rumen. Appl Microbiol Biotechnol 58:663–668. 23. Samuelov NS, Lamed R, Lowe SE, Zeikus JG. 1991. Influence of CO2-HCO3- levels and pH on growth, succinate production, and enzyme activities of Anaerobiospirillum succiniciproducens. Appl Environ Microbiol 57:3013–3019. 24. Meynial‐Salles I, Dorotyn S, Soucaille P. 2008. A new process for the continuous production of succinic acid from glucose at high yield, titer, and productivity. Biotechnol Bioengin 99:129-135. 33 25. Secchi C, Cantarelli VV, Pereira Fde S, Wolf HH, Brodt TC, Amaro MC, Inamine E. 2005. Fatal bacteremia due to Anaerobiospirillum succiniciproducens: first description in Brazil. Braz J Infect Dis 9:169-172. 26. van der Werf MJ, Guettler MV, Jain MK, Zeikus JG. 1997. Environmental and physiological factors affecting the succinate product ratio during carbohydrate fermentation by Actinobacillus sp. 130Z. Arch Microbiol 167:332–342. 27. McKinlay JB, Shachar-Hill Y, Zeikus JG, Vieille C. 2007. Determining Actinobacillus succinogenes metabolic pathways and fluxes by NMR and GC-MS analyses of 13Clabeled metabolic product isotopomers. Metab Eng 9:177–192. 28. McKinlay JB, Zeikus JG, Vieille C. 2005. Insights into Actinobacillus succinogenes fermentative metabolism in a chemically defined growth medium. Appl Environ Microbiol 71:6651-6656. 29. McKinlay JB, Vieille C. 2008. 13C-metabolic flux analysis of Actinobacillus succinogenes fermentative metabolism at different NaHCO3 and H2 concentrations. Metab Eng 10:55–68. 30. McKinlay JB, Laivenieks M, Schindler BD, Mckinlay AA, Siddaramappa S, Challacombe JF, Lowry SR, Clum A, Lapidus AL, Burkhart KB, Harkins V, Vieille C. 2010. A genomic perspective on the potential of Actinobacillus succinogenes for industrial succinate production. BMC Genomics 11:680. 31. Joshi RJ, Schindler BD, McPherson NR, Tiwari K, Vieille C. 2014. Development of a markerless knockout method for Actinobacillus succinogenes. Appl Environ Microbiol 80:3053-3061. 32. Carvalho M, Matos M, Roca C, Reis MAM. 2014. Succinic acid production from glycerol by Actinobacillus succinogenes using dimethylsulfoxide as electron acceptor. N Biotechnol 31:133-139. 33. Van Heerden CD, Nicol W. 2013. Continuous succinic acid fermentation by Actinobacillus succinogenes. Biochem Engin J 73:5-11. 34. Bradfield MFA, Nicol W. 2014. Continuous succinic acid production by Actinobacillus succinogenes in a biofilm reactor: steady-state metabolic flux variation. Biochem Engin J 85:1-7. 35. Guettler M, Hanchar R, Kleff S, Jadhav S. 2015. Recombinant microorganisms for producing organic acids. 36. Vlysidis A, Binns M, Webb C, Theodoropoulos C. 2009. Utilisation of glycerol to platform chemicals within the biorefinery concept: a case for succinate production. Chem Eng Trans 18:537-542. 34 37. Vlysidis A, Binns M, Webb C, Theodoropoulos C. 2011. Glycerol utilisation for the production of chemicals: conversion to succinic acid, a combined experimental and computational study. Biochem Eng J 58:1-11. 38. Pateraki C, Almqvist H, Ladakis D, Liden G, Koutinas AA, Vlysidis A. 2016. Modelling succinic acid fermentation using a xylose based substrate. Biochem Eng J 114:26-41. 39. Salvachúa D, Mohagheghi A, Smith H, Bradfield MFA, Nicol W, Black BA, Biddy MJ, Dowe N, Beckham GT. 2016. Succinic acid production on xylose-enriched biorefinery streams by Actinobacillus succinogenes in batch fermentation. Biotechnol Biofuels 9:28. 40. Guarnieri MT, Chou YC, Salvachua D, Mohagheghi A, St John PC, Peterson DJ, Bomble YJ, Beckham GT. 2017. Metabolic Engineering of Actinobacillus succinogenes Provides Insights into Succinic Acid Biosynthesis. Appl Environ Microbiol 83. 41. Zheng P, Dong J-J, Sun Z-H, Ni Y, Fang L. 2009. Fermentative production of succinic acid from straw hydrolysate by Actinobacillus succinogenes. Bioresour Technol 100:2425-2429. 42. Li F, Wang Y, Gong K, Wang Q, Liang Q, Qi Q. 2014. Constitutive expression of RyhB regulates the heme biosynthesis pathway and increases the 5-aminolevulinic acid accumulation in Escherichia coli. FEMS Microbiol Lett 350:209-215. 43. Zheng P, Fang L, Xu Y, Dong J-J, Ni Y, Sun Z-H. 2010. Succinic acid production from corn stover by simultaneous saccharification and fermentation using Actinobacillus succinogenes. Bioresour Technol 101:7889-7894. 44. Wan C, Li Y, Shahbazi A, Xiu S. 2008. Succinic acid production from cheese whey using Actinobacillus succinogenes 130 Z. Appl Biochem Biotech 145:111-119. 45. Lee SJ, Song H, Lee SY. 2006. Genome-based metabolic engineering of Mannheimia succiniciproducens for succinic acid production. Appl Environ Microbiol 72:1939–1948. 46. Choi S, Song H, Lim SW, Kim TY, Ahn JH, Lee JW, Lee M-H, Lee SY. 2016. Highly selective production of succinic acid by metabolically engineered Mannheimia succiniciproducens and its efficient purification. Biotechnol Bioeng 113:2168-2177. 47. Lee JW, Yi J, Kim TY, Choi S, Ahn JH, Song H, Lee M-H, Lee SY. 2016. Homosuccinic acid production by metabolically engineered Mannheimia succiniciproducens. Metab Eng 38:409-417. 48. Pettigrew DW, Liu WZ, Holmes C, Meadow ND, Roseman S. 1996. A single amino acid change in Escherichia coli glycerol kinase abolishes glucose control of glycerol utilization in vivo. J Bacteriol 178:2846-2852. 35 49. Lee KS. 2004. Anaerobic hydrogen production with an efficient carrier-induced granular sludge bed bioreactor. Biotechnol Bioeng 87:648-657. 50. Scholten E, Dägele D. 2008. Succinic acid production by a newly isolated bacterium. Biotechnol Lett 30:2143-2146. 51. Scholten E, Renz T, Thomas J. 2009. Continuous cultivation approach for fermentative succinic acid production from crude glycerol by Basfia succiniciproducens DD1. Biotechnol Lett 31:1947-1951. 52. Becker J, Reinefeld J, Stellmacher R, Schäfer R, Lange A, Meyer H, Lalk M, Zelder O, von Abendroth G, Schröder H. 2013. Systems-wide analysis and engineering of metabolic pathway fluxes in bio‐succinate producing Basfia succiniciproducens. Biotechnol Bioeng 110:3013-3023. 53. Chatterjee R, Millard CS, Champion K, Clark DP, Donnelly MI. 2001. Mutation of the ptsG gene results in increased production of succinate in fermentation of glucose by Escherichia coli. Appl Environ Microbiol 67:148–154. 54. Thakker C, Martínez I, San K-Y, Bennett GN. 2012. Succinate production in Escherichia coli. Biotechnol J 7:213-224. 55. Sawers RG, Clark DP. 2004. Fermentative pyruvate and acetyl-coenzyme A metabolism. In Bock A (ed), Escherichia coli and Salmonella: Cellular and Molecular Biology. ASM Press. 56. Millard CS, Chao YP, Liao JC, Donnelly MI. 1996. Enhanced production of succinic acid by overexpression of phosphoenolpyruvate carboxylase in Escherichia coli. Appl Environ Microbiol 62:1808–1810. 57. Gokarn RR, Eiteman MA, Altman E. 2000. Metabolic analysis of Escherichia coli in the presence and absence of the carboxylating enzymes phosphoenolpyruvate carboxylase and pyruvate carboxylase. Appl Environ Microbiol 66:1844–1850. 58. Chatterjee T, Datta AG. 1973. Anaerobic formation of succinate from glucose and bicarbonate in resting cells of Leishmania donovani. Exp Parasitol 33:138-146. 59. Vemuri GN, Eiteman MA, Altman E. 2002. Succinate production in dual-phase Escherichia coli fermentations depends on the time of transition from aerobic to anaerobic conditions. J Ind Microbiol Biotechnol 28:325-332. 60. Kim P, Laivenieks M, McKinlay J, Vieille C, Zeikus JG. 2004. Construction of a shuttle vector for the overexpression of recombinant proteins in Actinobacillus succinogenes. Plasmid 51:108–115. 36 61. Jantama K, Haupt MJ, Svoronos SA, Zhang XL, Moore JC, Shanmugam KT, Ingram LO. 2008. Combining metabolic engineering and metabolic evolution to develop nonrecombinant strains of Escherichia coli C that produce succinate and malate. Biotechnol Bioeng 99:1140-1153. 62. Jantama K, Zhang XL, Moore JC, Shanmugam KT, Svoronos SA, Ingram LO. 2008. Eliminating side products and increasing succinate yields in engineered strains of Escherichia coli C. Biotechnol Bioeng 101:881-893. 63. Zhang XL, Jantama K, Moore JC, Jarboe LR, Shanmugam KT, Ingram LO. 2009. Metabolic evolution of energy-conserving pathways for succinate production in Escherichia coli. Proc Natl Acad Sci USA 106:20180-20185. 64. Zhang XL, Jantama K, Shanmugam KT, Ingram LO. 2009. Reengineering Escherichia coli for succinate production in mineral salts medium. Appl Environ Microbiol 75:7807-7813. 65. Zhang X, Shanmugam KT, Ingram LO. 2010. Fermentation of glycerol to succinate by metabolically engineered strains of Escherichia coli. Appl Environ Microbiol 76:23972401. 66. Inui M, Murakami S, Okino S, Kawaguchi H, Vertès AA, Yukawa H. 2004. Metabolic analysis of Corynebacterium glutamicum during lactate and succinate productions under oxygen deprivation conditions. J Mol Microbiol Biotechnol Bioeng 7:182-196. 67. Okino S, Inui M, Yukawa H. 2005. Production of organic acids by Corynebacterium glutamicum under oxygen deprivation. Appl Microbiol Biotechnol 68:475-480. 68. Okino S, Noburyu R, Suda M, Jojima T, Inui M, Yukawa H. 2008. An efficient succinic acid production process in a metabolically engineered Corynebacterium glutamicum strain. Appl Microbiol Biotechnol 81:459-464. 69. Litsanov B, Brocker M, Bott M. 2012. Toward homosuccinate fermentation: metabolic engineering of Corynebacterium glutamicum for anaerobic production of succinate from glucose and formate. Appl Environ Microbiol 78:3325-3337. 70. Zhu N, Xia H, Yang J, Zhao X, Chen T. 2014. Improved succinate production in Corynebacterium glutamicum by engineering glyoxylate pathway and succinate export system. Biotechnol Lett 36:553-560. 71. Litsanov B, Kabus A, Brocker M, Bott M. 2012. Efficient aerobic succinate production from glucose in minimal medium with Corynebacterium glutamicum. Microb Biotechnol 5:116-128. 37 72. Litsanov B, Brocker M, Bott M. 2013. Glycerol as a substrate for aerobic succinate production in minimal medium with Corynebacterium glutamicum. Microb Biotechnol 6:189-195. 73. Zhu N, Xia H, Wang Z, Zhao X, Chen T. 2013. Engineering of acetate recycling and citrate synthase to improve aerobic succinate production in Corynebacterium glutamicum. PLoS One 8:e60659. 74. Chung SC, Park JS, Yun J, Park JH. 2017. Improvement of succinate production by release of end-product inhibition in Corynebacterium glutamicum. Metab Eng 40:157164. 75. Arikawa Y, Kobayashi M, Kodaira R, Shimosaka M, Muratsubaki H, Enomoto K, Okazaki M. 1999. Isolation of sake yeast strains possessing various levels of succinateand/or malate-producing abilities by gene disruption or mutation. J Biosci Bioeng 87:333-339. 76. Kubo Y, Takagi H, Nakamori S. 2000. Effect of gene disruption of succinate dehydrogenase on succinate production in a sake yeast strain. J Biosci Bioeng 90:619624. 77. Raab AM, Gebhardt G, Bolotina N, Weuster-Botz D, Lang C. 2010. Metabolic engineering of Saccharomyces cerevisiae for the biotechnological production of succinic acid. Metab Eng 12:518-525. 78. Ito Y, Hirasawa T, Shimizu H. 2014. Metabolic engineering of Saccharomyces cerevisiae to improve succinic acid production based on metabolic profiling. Biosci Biotechnol Biochem 78:151-159. 79. Yan D, Wang C, Zhou J, Liu Y, Yang M, Xing J. 2014. Construction of reductive pathway in Saccharomyces cerevisiae for effective succinic acid fermentation at low pH value. Bioresour Technol 156:232-239. 80. Scholten E, Haefner S, Schröder H. 2014. Bacterial cells having a glyoxylate shunt for the manufacture of succinic acid. US Patent 8,877,466 B2. 81. Schroder H, Haefner S, Von Abendroth G, Hollmann R, Raddatz A, Ernst H, Gurski H. 2014. Microbial succinic acid producers and purification of succinic acid. 82. Janausch IG, Zientz E, Tran QH, Kroger A, Unden G. 2002. C4-dicarboxylate carriers and sensors in bacteria. Biochim Biophys Acta 1553:39-56. 83. Zientz E, Six S, Unden G. 1996. Identification of a third secondary carrier (DcuC) for anaerobic C4-dicarboxylate transport in Escherichia coli: roles of the three Dcu carriers in uptake and exchange. J Bacteriol 178:7241-7247. 38 84. Karinou E, Compton EL, Morel M, Javelle A. 2013. The Escherichia coli SLC26 homologue YchM (DauA) is a C4-dicarboxylic acid transporter. Mol Microbiol 87:623640. 85. Six S, Andrews SC, Unden G, Guest JR. 1994. Escherichia coli possesses two homologous anaerobic C4-dicarboxylate membrane transporters (DcuA and DcuB) distinct from the aerobic dicarboxylate transport system (Dct). J Bacteriol 176:64706478. 86. Davies SJ, Golby P, Omrani D, Broad SA, Harrington VL, Guest JR, Kelly DJ, Andrews SC. 1999. Inactivation and regulation of the aerobic C4-dicarboxylate transport (dctA) gene of Escherichia coli. J Bacteriol 181:5624-5635. 87. Youn JW, Jolkver E, Kramer R, Marin K, Wendisch VF. 2009. Characterization of the dicarboxylate transporter DctA in Corynebacterium glutamicum. J Bacteriol 191:5480-5488. 88. Fukui K, Nanatani K, Hara Y, Yamakami S, Yahagi D, Chinen A, Tokura M, Abe K. 2017. Escherichia coli yjjPB genes encode a succinate transporter important for succinate production. Biosci Biotechnol Biochem 81:1837-1844. 89. Chen J, Zhu X, Tan Z, Xu H, Tang J, Xiao D, Zhang X. 2014. Activating C4dicarboxylate transporters DcuB and DcuC for improving succinate production. Appl Microbiol Biotechnol 98:2197-2205. 90. Kelly DJ, Thomas GH. 2001. The tripartite ATP-independent periplasmic (TRAP) transporters of bacteria and archaea. FEMS Microbiol Rev 25:405-424. 91. Mulligan C, Fischer M, Thomas GH. 2011. Tripartite ATP-independent periplasmic (TRAP) transporters in bacteria and archaea. FEMS Microbiol Rev 35:68-86. 92. Pos KM, Dimroth P, Bott M. 1998. The Escherichia coli citrate carrier CitT: a member of a novel eubacterial transporter family related to the 2-oxoglutarate/malate translocator from spinach chloroplasts. J Bacteriol 180:4160-4165. 93. Huhn S, Jolkver E, Kramer R, Marin K. 2011. Identification of the membrane protein SucE and its role in succinate transport in Corynebacterium glutamicum. Appl Microbiol Biotechnol 89:327-335. 94. Fukui K, Koseki C, Yamamoto Y, Nakamura J, Sasahara A, Yuji R, Hashiguchi K, Usuda Y, Matsui K, Kojima H, Abe K. 2011. Identification of succinate exporter in Corynebacterium glutamicum and its physiological roles under anaerobic conditions. J Biotechnol 154:25-34. 39 95. Rhie MN, Yoon HE, Oh HY, Zedler S, Unden G, Kim OB. 2014. A Na+-coupled C4dicarboxylate transporter (Asuc_0304) and aerobic growth of Actinobacillus succinogenes on C4-dicarboxylates. Microbiology 160:1533-1544. 96. Na D, Yoo SM, Chung H, Park H, Park JH, Lee SY. 2013. Metabolic engineering of Escherichia coli using synthetic small regulatory RNAs. Nat Biotechnol 31:170-174. 97. Kim JYH, Cha HJ. 2003. Down-regulation of acetate pathway through antisense strategy in Escherichia coli: improved foreign protein production. Biotechnol Bioeng 83:841-853. 98. Nakashima N, Tamura T, Good L. 2006. Paired termini stabilize antisense RNAs and enhance conditional gene silencing in Escherichia coli. Nucleic Acids Res 34:e138. 99. Negrete A, Majdalani N, Phue J-N, Shiloach J. 2013. Reducing acetate excretion from E. coli K-12 by over-expressing the small RNA SgrS. N Biotechnol 30:269-273. 100. Coleman J, Green PJ, Inouye M. 1984. The use of RNAs complementary to specific mRNAs to regulate the expression of individual bacterial genes. Cell 37:429-436. 101. Kang Z, Wang X, Li Y, Wang Q, Qi Q. 2012. Small RNA RyhB as a potential tool used for metabolic engineering in Escherichia coli. Biotechnol Lett 34:527-531. 102. Ding W, Weng H, Du G, Chen J, Kang Z. 2017. 5-Aminolevulinic acid production from inexpensive glucose by engineering the C4 pathway in Escherichia coli. J Ind Microbiol Biotechnol 44:1127-1135. 103. Desai RP, Papoutsakis ET. 1999. Antisense RNA strategies for metabolic engineering of Clostridium acetobutylicum. Appl Environ Microbiol 65:936-945. 104. Boudry P, Gracia C, Monot M, Caillet J, Saujet L, Hajnsdorf E, Dupuy B, MartinVerstraete I, Soutourina O. 2014. Pleiotropic role of the RNA chaperone protein Hfq in the human pathogen Clostridium difficile. J Bacteriol 196:3234-3248. 105. Liu Y, Zhu Y, Li J, Shin H-D, Chen RR, Du G, Liu L, Chen J. 2014. Modular pathway engineering of Bacillus subtilis for improved N-acetylglucosamine production. Metab Eng 23:42-52. 106. Cho C, Lee SY. 2016. Efficient gene knockdown in Clostridium acetobutylicum by synthetic small regulatory RNAs. Biotechnol Bioeng 114:374-383. 107. Subashchandrabose S, Leveque RM, Kirkwood RN, Kiupel M, Mulks MH. 2013. The RNA chaperone Hfq promotes fitness of Actinobacillus pleuropneumoniae during porcine pleuropneumonia. Infect Immun 81:2952-2961. 40 108. Amarasinghe JJ, Connell TD, Scannapieco FA, Haase EM. 2012. Novel iron‐ regulated and Fur‐regulated small regulatory RNAs in Aggregatibacter actinomycetemcomitans. Mol Oral Microbiol 27:327-349. 109. Baddal B, Muzzi A, Censini S, Calogero RA, Torricelli G, Guidotti S, Taddei AR, Covacci A, Pizza M, Rappuoli R. 2015. Dual RNA-seq of nontypeable Haemophilus influenzae and host cell transcriptomes reveals novel insights into host-pathogen cross talk. mBio 6:e01765-15. 110. Santana EA, Harrison A, Zhang X, Baker BD, Kelly BJ, White P, Liu Y, Munson Jr RS. 2014. HrrF is the Fur-regulated small RNA in nontypeable Haemophilus influenzae. PloS one 9:e105644. 111. Rossi CC, Bossé JT, Li Y, Witney AA, Gould KA, Langford PR, Bazzolli DMS. 2016. A computational strategy for the search of regulatory small RNAs in Actinobacillus pleuropneumoniae. RNA 22:1373-1385. 41 Chapter 2 Identification of succinate transporters in Actinobacillus succinogenes 42 2.1 Abstract Succinate, listed as one of the top 12 building block chemicals by the U.S. Department of Energy has been produced from maleic anhydride, a petrochemical until recently. A few companies, including BioAmber, Myriant, and Reverdia, have started producing industrial scale amounts of bio-based succinate using yeast and E. coli strains. A. succinogenes, along with other succinate producing bacteria have been studied and engineered with the aim of increasing succinate production. Succinate exporters have been studied in very few of these bacteria, especially with the aim of using them as targets in engineering for increased succinate production. E. coli and C. glutamicum are the only strains where succinate transporters have been studied. Several succinate exporter candidates have been overexpressed in E. coli and in C. glutamicum and caused an increase in succinate efflux. So far, A. succinogenes has not been studied for succinate exporters. In this study, we sought to identify succinate transporters using a combined proteomics and transcriptomics approach. The proteins with the most hits in our proteomics studies— Asuc_1999, Asuc_0142, Asuc_2058 and Asuc_1990-91 were selected for overexpression in the wild-type strain. Since we are limited by our only expression vector which harbors the native, strong, and constitutive promoter ppckA, we developed a library of expression vectors with promoters weaker than ppckA by truncating ppckA and searching for additional promoter candidates from our RNA sequencing analysis. We looked for candidate genes that were expressed at a similar level across growth conditions tested and that had several fold lower transcript levels than pckA. Finally, we overexpressed our candidate succinate exporters under the control of promoter’s ppckA-92 and ppckA-103, and observed an increase in succinate production in all the strains tested as compared to the control strain. The highest increase was seen when Asuc_1999 43 was expressed under the control of ppckA-103. To the best of our knowledge this is the first study focusing on identifying succinate exporters in A. succinogenes. This is also the first report of a promoter library being constructed for A. succinogenes, further increasing the tunability of gene expression. 2.2 Introduction Succinate is a dicarboxylate molecule with many possible applications. It can be used as a precursor to make products such as nylons, resins, paints, cosmetics, flavors, and adhesives (1). The conventional route to produce succinate has been by hydrogenation of maleic anhydride, a petroleum derivative, in the presence of a metal catalyst (2). Bio-based succinate production has gained momentum in recent years, though, as awareness about the benefits of renewable products and downsides of burning fossil fuels have increased. A number of companies, such as BioAmber, Myriant, and Succinity, have started selling bio-based succinate produced by engineered microorganisms (e.g., Escherichia coli, Basfia succiniciproducens, and yeast). Biobased succinate is produced naturally by anaerobes such as Actinobacillus succinogenes and Mannheimia succiniproducens via the reductive pathway of the tricarboxylic acid cycle. Industrial workhorse microorganisms such as E. coli, Saccharomyces cerevisiae, and Corynebacterium glutamicum also have been or are being engineered to produce succinate (2-5). One of the major hurdles still lies in making bio-based succinate as cost–effective as conventionally produced succinate. A. succinogenes is one of the best natural succinate producers, yet our knowledge of its uptake and efflux mechanisms remains limited. For example, the system or systems involved in glucose uptake as well as the one or ones involved in succinate export remain unknown, and the succinate export system has not yet been targeted for increasing succinate production. 44 Several succinate transporters have been identified and characterized in E. coli, but most studies have focused on uptake and antiport transporters. The dicarboxylate uptake (Dcu) carriers DcuA, B, and C are the best-characterized succinate transporters. DcuA and DcuB are involved in fumarate: succinate antiport during anaerobic respiration on fumarate. They are also capable of uptake and efflux of C4-dicarboxylates (6, 7). Zeintz et al. (8) suggested that, even though DcuC can carry out the same transport functions as DcuA and DcuB, its major function likely is succinate export. E. coli dcuD shows similarity to dcuC, but DcuD mutants show no effect on E. coli growth under a number of conditions, and do not affect succinate uptake or efflux (9). DauA is the sole transporter for aerobic succinate uptake at acidic pH (10). Another transporter, CitT, catalyzes citrate/succinate antiport during citrate fermentation (11). DctA, a H+/C4-dicarboxylate symporter is responsible for C4-dicarboxylate uptake in aerobic conditions (6, 12) at neutral pHs. In one report, a quintuple dctA dcuA dcuB dcuC citT E. coli mutant did not show any deficiency in succinate efflux as compared to the wild-type strain in glucose-grown anaerobic cultures (13). In another report, while deleting dcuA and dcuD did not have any effect on succinate efflux, individual dcuB and dcuC deletion mutants showed decreases in succinate titers of 15% and 11%, respectively (14). A double dcuBC deletion resulted in a 90% decrease in succinate titer, and modulating expression of both genes increased the succinate titer by 34% (14). Thus no clear picture of the involvement of Dcu transporters in succinate efflux is available. DauA is the sole characterized E. coli succinate transporter that has not been tested for succinate efflux during glucose fermentation. Two studies using different approaches identified the same succinate exporter, SucE, of the aspartate:alanine exchanger family, in Corynebacterium glutamicum (15, 16). This transporter functioned in anaerobic and microaerobic conditions and was responsible for about 45 30% of the succinate export in microaerobic conditions, indicating that SucE is not the sole succinate exporter in C. glutamicum. In a recent study aimed at identifying the E. coli anaerobic succinate exporter, C. glutamicum strain AJ110655 ΔsucE, deficient in succinate export, was partially rescued for succinate export by expressing the E. coli yjjPB genes (17). Additionally, an E. coli yjjPB knockout mutant produced 70% less succinate than the parental strain, indicating that YjjPB may be one of the main E. coli anaerobic succinate exporters. There has been no report on whether or not the sole E. coli SucE homolog participates in succinate export. In this study, we conducted transcriptomic studies of A. succinogenes grown on multiple carbon sources as well as proteomics studies on cytoplasmic membrane-enriched fractions of A. succinogenes grown anaerobically on glucose, xylose, and fructose (conditions favoring succinate production) or on glycerol-nitrate (less succinate is produced). Of the putative C4 dicarboxylate transporters identified in the annotated genome, locus tags Asuc_1999, Asuc_1990-1991, Asuc_0142, and Asuc_2058 had the most hits in our proteomics and transcriptomics results, and we focused on these four transporters for further analysis. Currently, to overexpress A. succinogenes genes we only have one strong constitutive native promoter (ppckA, (18) at our disposal, which is not optimal for the expression of membrane proteins. We constructed a library of promoters of varying strengths to overcome this problem and modulate the expression of these transporters. We show that overexpression of any of these four putative C4-dicarboxylate transporters significantly increases succinate production. 46 2.3 Materials and Methods 2.3.1 Strains, media, and culture conditions. E. coli strains (Table 2.1) were cultivated in lysogeny broth (LB) and on LB agar plates and were supplemented with 100 µg mL-1 ampicillin or 50 µg mL-1 kanamycin when required for plasmid maintenance (19). A. succinogenes type strain 130Z was obtained from the American Type Culture Collection. A. succinogenes liquid cultures were grown in AM3 defined medium (20), Medium B (21), or Bacto brain heart infusion medium (BHI; Becton, Dickinson and Co., Franklin Lakes, NJ). Mutant strains of A. succinogenes were grown on AM2-isocitrate or AM16isocitrate after natural transformations (21). Liquid cultures were grown anaerobically in 28-mL anaerobic tubes flushed with N2 at 37°C with shaking at 250 rpm. Optical densities were measured at 660 nm using a Spectronic 20 (Bausch and Lomb, Rochester, NY) or a DU650 spectrophotometer (Beckman, Fullerton, CA). To isolate A. succinogenes colonies, strains were grown on LB plates containing 10 g L-1 glucose (22) and 10 µg mL-1 kanamycin (Km), 600 mg L-1 polyvinyl alcohol (PVA) and 10 mg L-1 CaCl2. Agar plate cultures of A. succinogenes strains were grown under a CO2-enriched atmosphere for 24-48 h. All strains used in this study are listed in Table 2.1. 2.3.2 Plasmids, DNA manipulations, and electroporations. PCR products were cloned into pCR2.1 using the TA TOPO cloning kit (Invitrogen, Carlsbad, CA) and transformed into One Shot TOP10 chemically competent cells (Invitrogen) as per manufacturer’s protocols. Shuttle plasmid pLGZ920 was used to express native or foreign genes in A. succinogenes (18). Electrocompetent A. succinogenes cells, prepared as described (21), were transformed with plasmid constructs and spread on LB Glucose PVA CaCl2 Km plates. 47 DNA manipulations were carried out according to Ausubel et al. (19). PCR products were amplified using Phusion High-Fidelity DNA polymerase (New England Biolabs, Ipswich, MA). Restriction enzymes were from New England Biolabs. Plasmids and DNA fragments were purified using the Wizard SV miniprep kit and Wizard SV Gel and PCR Clean-Up system, respectively, (Promega, Madison, WI). Primers used in this work were synthesized by Integrated DNA Technologies (Coralville, IA) and are listed in Table S2.1 in supplementary material. DNA sequencing was performed by GENEWIZ, Inc. (South Plainfield, NJ). Colonies were screened by colony PCR using Taq polymerase. Table 2.1 Strains and plasmids used in this study Description Source DH5α F- φ80lacZΔM15 Δ(lacZYA-argF) U169 recA1 endA1 hsdR17 Laboratory collection TOP10 (rk-, mk+) phoA supE44 λ- thi-1 gyrA96 relA1 F- mcrA Δ(mrrhsdRMS-mcrBC) φ80lacZΔM15 ΔlacΧ74 recA1 araD139 Δ(ara-leu) 7697 galU galK rpsL (StrR) endA1 nupG λ- Invitrogen 130Z (ATCC55618) Wild-type strain ATCC ΔlacZ::icd 130Z derivative, contains the AscI-FRT-ppckA-icd-FRT-AscI cassette in the lacZ deletion This study ΔlacZ 130Z derivative, lacZ deletion, contains one FRT site This study pCR2.1 TOPO AmpR, KmR, lacZα, cloning vector Invitrogen pCR2.1-ΔlacZ pCR2.1 derivative, lacZ deletion (2.1-kb fusion product of frdup and frdCD) (21) pCR2.1-icd pCR2.1 derivative, E. coli icd under control of A. succinogenes ppckA, flanked by FRT repeats and AscI restriction sites (21) pCV933 pLGZ920 derivative, S. cerevisiae flp under control of ppckA (21) pLGZ920-lacZ pLGZ920 derivative containing lacZ under ppckA This study pLGZ920-ppckA-92-lacZ pLGZ920 derivative containing lacZ under ppckA-92 This study E. coli A. succinogenes Plasmids 48 Table 2.1 (cont’d) pLGZ920-ppckA-103-lacZ pLGZ920 derivative containing lacZ under ppckA-103 This study pLGZ920-ppckA-134-lacZ pLGZ920 derivative containing lacZ under ppckA-134 This study pLGZ920-ppckA-164-lacZ pLGZ920 derivative containing lacZ underppckA-164 This study pLGZ920*-pAsuc_0701 pLGZ920 derivative with the HindIII site 3’ of ColE1 ori replaced with This study BamHI and ppckA-0701 pLGZ920*-pAsuc_0701-lacZ pLGZ920* derivative containing lacZ under pAsuc_0701 This study pLGZ920*-pAsuc_2109-lacZ pLGZ920* derivative containing lacZ under pAsuc_2109 This study pLGZ920*-pAsuc_0391-lacZ pLGZ920* derivative containing lacZ under pAsuc_0391 This study pLGZ920*-pAsuc_0289-lacZ pLGZ920* derivative containing lacZ under pAsuc_0289 This study 2.3.3 Preparation of A. succinogenes cytoplasmic membrane-enriched fractions for proteomics analysis. A. succinogenes cultures in AM3 glucose, fructose, and xylose were grown in 300-mL volumes and harvested in the middle of the exponential phase at an OD660 of 0.8 to 0.9. Because A. succinogenes does not grow fermentatively on glycerol, glycerol cultures were grown by nitrate respiration. The glycerol-nitrate cultures were grown in 600-mL volumes and harvested at an OD660 no higher than 0.4 to 0.5 to avoid succinate accumulation after nitrate depletion (Schindler et al, 2014). Cytoplasmic membrane-enriched fractions were prepared using a method adapted from (23). Cells were harvested by centrifugation at 12,000 × g for 10 min and washed twice with 100 mM Tris-HCl (pH 8.0). Bacterial pellets were resuspended in 15 mL of buffer A (100 mM Tris-HCl [pH 8.0] containing 20% sucrose and Complete Mini, EDTA-free, protease inhibitor cocktail [Sigma-Aldrich]). ReadyLyse lysozyme (EpiCentre, Madison, WI) was added 49 at a final concentration of 104 U mL-1, and bacterial suspensions were stirred on a magnetic plate at the lowest setting for 8 min. Ethylenediaminetetraacetic acid (EDTA, 1.5 mL of 100 mM, pH 8.0) (1:10 v/v EDTA:cells) was added drop by drop over 2.5 min to avoid lysis, and suspensions were stirred for 8 more min. The outer membrane and periplasmic fractions were separated from the spheroplasts by centrifugation at 12,000 × g for 10 min. The pellets, corresponding to the spheroplast-enriched fractions, were washed once with 15 mL buffer A, and resuspend in 15 mL buffer A containing 225,000 U DNase I and 0.2 mg RNase A. The spheroplasts were lysed by passing twice though a French press at 1,400 psi (high setting). The intact spheroplasts were removed by centrifugation at 12,000 × g for 10 min. The supernatants were then ultracentrifuged at 150,000 × g for 2 h at 4°C. The pellets were resuspended in buffer A and ultracentrifuged again to wash off the cytoplasmic proteins. The cytoplasmic membrane-enriched fractions were finally resuspended in 100 mM Tris-HCl (pH 8.0) containing 1 mM EDTA and protease inhibitors to a concentration of 4 mg mL-1. 2.3.4 Proteomic analysis. To submit the cytoplasmic membrane-enriched fractions for proteomic analysis, 50 µl of extracts (i.e., 200 µg protein) were loaded on sodium dodecyl sulfate-polyacrylamide gels (miniPROTEAN TGX gel, 10% acrylamide, Bio-Rad, Hercules, CA) and subjected to electrophoresis until the samples had entered the stacking gel. The proteins were visualized by Coomassie blue staining. Gel bands were digested in-gel according to Shevchenko et al.1 with modifications. Briefly, gel bands were dehydrated using 100% acetonitrile and incubated with 10 mM dithiothreitol in 100 mM ammonium bicarbonate (pH 8.0) at 56 °C for 45 min, dehydrated again and incubated in the dark with 50 mM iodoacetamide in 100 mM ammonium bicarbonate for 20 50 min. Gel bands were then washed with 100 mM ammonium bicarbonate and dehydrated again. Sequencing-grade modified trypsin (50 μL at 0.01 μg μL-1 in 50 mM ammonium bicarbonate) was added to each gel band to completely submerge the band. Bands were then incubated at 37 °C overnight. Peptides were extracted from the gel by water bath sonication in a solution of 60% acetonitrile/1% trichloroacetic acid and vacuum-dried to ~2μL. Peptides were then re-suspended in 2% acetonitrile/0.1% trifluoroacetic acid to 25 μL. Five μL were automatically injected by a Thermo (www.thermo.com) EASYnLC 1000 onto a Thermo Acclaim PepMap RSLC 0.075 mm x 250 mm C18 column and eluted over 185 min with a 174-min gradient of 5% B to 25% B. The gradient was then raised to 100% B in 1 min and held at 100% B for the duration of the run (Buffer A = 99.9% water/0.1% formic acid, Buffer B = 99.9% acetonitrile/0.1% formic Acid). Eluted peptides were sprayed into a Q-Exactive mass spectrometer (Thermo Fisher Scientific, Waltham, MA) using a FlexSpray spray ion source. Survey scans were taken in the Orbi trap (35,000 resolution, determined at m/z 200) and the top ten ions in each survey scan were subjected to automatic higher energy collision induced dissociation with fragment spectra acquired at 17,500 resolution. The resulting MS/MS spectra were converted to peak lists using Mascot Distiller, v2.5.1 (www.matrixscience.com) and searched against a protein sequence database containing entries for A. succinogenes (downloaded on May 22, 2015 from NCBI, www.ncbi.nlm.nih.gov) and appended with common laboratory contaminants (downloaded from www.thegpm.org, cRAP project) using the Mascot searching algorithm, v 2.5. The following parameters were used: allow up to two missed tryptic sites, fixed modification of carbamidomethyl cysteines, variable modification of methionine oxidation and of asparagine and glutamine deamidation; +/- 5 ppm 51 peptide tolerance, 0.3 Da MS/MS tolerance, and false discovery rate (FDR) calculated using randomized database search. The Mascot output was then analyzed using Scaffold, v4.4.3 (www.proteomesoftware.com) to probabilistically validate protein identifications. Assignments validated using the Scaffold 1% FDR confidence filter were considered true. NSAF values (24) were calculated within Scaffold for protein quantitation. 2.3.5 Total RNA purification and RNA sequencing Total RNA was purified from A. succinogenes cultures and quality-controlled as described (25). In short, cultures harvested in the middle of the exponential phase were immediately mixed with one volume of -20 °C methanol and stored at -20 °C for at least 30 min to quench all enzymatic activity. After centrifugation, RNA was purified from the cell pellets with the QIAGEN RNeasy Midi Kit (QIAGEN, Hilden, Germany), including the optional on-column DNase digestion. RNA samples were quantified on a Qubit Fluorometer (Life Technologies, Grand Island, NY), and RNA quality was validated visually on a 1% agarose gel. RNA was mixed 50% v/v with formaldehyde before loading on the gel. Total RNA sequencing was performed by the Michigan State University Research Technology Support Facility as described (25). In short, libraries were prepared using the Illumina TruSeq Stranded Total RNA Library Preparation kit and rRNAs were depleted with the Illumina Ribo-Zero Gram-Negative Bacteria kit. Libraries were validated and quantified using the Qubit dsDNA assay, Caliper LabChipGX, and Kapa Illumina Library Quantification qPCR kits. The sequenced libraries (three biological replicates each) were from A. succinogenes 130Z fermenting glucose, xylose, and fructose, as well as A. succinogenes respiring glucose with nitrate as the electron acceptor, glycerol with nitrate as the electron acceptor, and glycerol in 52 microaerobic conditions. Pools of 12 libraries were loaded on separate lanes of an Illumina HiSeq 2500 High Output (v4) flow cell and sequenced in 1x50bp single end format using HiSeq SBS reagents. Base calling was done by Illumina Real Time Analysis v1.18.64. Output was demultiplexed and converted to FastQ format using Illumina Bcl2fastq v1.8.4. Sequencing results were analyzed using the SPARTA pipeline (26). 2.3.6 Identification of A. succinogenes’s pckA transcription start site. The transcriptional start site for pckA was identified with First Choice RNA ligasemediated rapid amplification of cDNA ends (RLM-RACE) kit (Ambion, Inc.) following the manufacturer’s instructions. DNase-treated total RNA (10 µg) extracted from microaerobic glycerol cultures (27) was treated with calf intestine phosphatase. The RNA was phenolchloroform extracted and then treated with tobacco acid pyrophosphatase (TAP) to remove the cap from full-length mRNAs. TAP-treated RNA was then used for 5’ RACE adapter ligation. The sample was reverse-transcribed and used as the template for nested PCR. Outer 5’RLMRACE PCR was carried out using the 5’ RACE outer primer and pckA-specific outer primer. The outer PCR product was used as the template for carrying out the inner 5’RLM-RACE PCR with the 5’RACE inner primer and pckA-specific inner primer. A-overhangs were added to the inner PCR product before cloning into pCR2.1. Colonies were screened by PCR using the inner 5’RLM-RACE PCR primers, and insert sequences were verified by sequencing. 2.3.7 Construction of truncated pckA promoters. Plasmid pLGZ920-lacZ was constructed to use lacZ as the reporter gene to test the strength of truncated pckA promoters. Primers CV624 and CV625 were used to amplify lacZ 53 from A. succinogenes 130Z genomic DNA. The PCR product was cloned into pLGZ920 under control of the ppckA promoter by Gibson Assembly (New England Biolabs, Ipswich, MA). The cloning reaction was transformed into cells provided with the kit. Colonies were screened by PCR with primers GZ1302 and GZ1303, and the insert sequence was verified by sequencing. Expression of lacZ from this plasmid construct was verified by β-galactosidase assay of recombinant strain ΔlacZ (pLGZ920-lacZ) (see below). The ColE1 origin of replication (ori) was amplified from pLGZ920 using primers CV469 and CV471. Primer pairs CV473-CV443, CV474-CV443, CV475-CV443, and CV476-CV443 were used to amplify truncated pckA promoters ppckA-92, ppckA-103, ppckA-134, and ppckA-164 with pLGZ920-lacZ as the template. The ColE1 ori and truncated pckA promoters were fused by PCR using primers CV469 and CV470. The four fusion PCR products were gel purified and cloned into the pLGZ920-lacZ HindIII and XbaI sites by Gibson assembly (New England Biolabs, Ipswich, MA) as described (Gibson et al., 2009). Note that restriction by HindIII removes the ColE1 ori initially present in pLGZ920. The cloning reactions were transformed into TOP10 cells. Colonies were screened by PCR using primers CV565 and CV568. Positive plasmids were verified by sequencing. The constructs were then transformed into strain ΔlacZ. 2.3.8 Construction of an A. succinogenes ΔlacZ mutant strain. The A. succinogenes ΔlacZ mutant strain was constructed as described (21). Briefly, the ΔlacZ::icd strain was built by natural transformation of A. succinogenes 130Z with a ΔlacZ::icd linear DNA fragment. Double recombinants were screened by colony PCR using primers CV376 and CV377. The icd marker was excised by the yeast flippase recombinase carried by plasmid pCV933. Transformants were screened for the loss of icd by colony PCR using primers CV376 54 and CV377. Plasmid pCV933 was cured from the ΔlacZ(pCV933) strain using acridine orange as described (21). Strain ΔlacZ was used as the host strain to test the promoter strength of the reporter promoter library constructs. 2.3.9 Preparation of crude extracts, β-galactosidase assays, and protein assays. Ten-mL cultures were grown in AM3 supplemented with 50 mM glucose and 150 mM NaHCO3. Cultures were harvested in exponential phase by centrifugation, washed, and resuspended in 0.5 mL of 50 mM sodium phosphate buffer (pH 7.0). Cells were lysed by sonication (Branson S-450A probe sonifier, Danbury, CT) with six 20-sec repetitions, 50% duty cycle, and a power level of 3. Cell extracts were centrifuged for 5 min (2,000 × g) at room temperature and the supernatant was stored on ice until used. Beta-galactosidase assays were conducted in 96-well plates in a PowerWave HT microplate reader at room temperature (BioTek Instruments Inc., Winooski, VT). Reactions (200 µl) contained 20 µl cell extract mixed with 150 µl of Z-buffer (60 mM Na2HPO4 , 40 mM NaH2PO4, 10 mM KCl, 1 mM MgSO4, 38 mM 2-mercaptoethanol) and 30 µL of 2-nitrophenyl β-Dgalactopyranoside (4 mg mL-1) in Z-buffer. Enzyme activity was calculated using the linear slope obtained from 2-nitrophenol production recorded at 420 nm. The extinction coefficient of 2nitrophenol was 4.8 mM-1 cm-1. Total cell protein was quantified using the Bio-Rad protein assay dye reagent using bovine serum albumin as the standard. Activities were reported as an average of three biological replicates. 2.3.10 Construction of a promoter library based on RNAseq results. An initial promoter-lacZ reporter plasmid was assembled in pLGZ920 using the promoter 55 of Asuc_0701 (pAsuc_0701). Other promoter-lacZ reporter plasmids were then constructed by substituting pAsuc_0701 with other promoters. First, ColE1 ori was amplified using primers CV615 and CV616 with pLGZ920 as the template. A. succinogenes 130Z genomic DNA was used as the template to amplify pAsuc_0701 with primers CV 617 and CV618. ColE1 ori and ppckA were removed from pLGZ920 by digestion with HindIII and XbaI. ColE1 ori and pAsuc_0701 were then cloned into pLGZ920’s HindIII and XbaI sites by Gibson Assembly to create plasmid pLGZ920*- pAsuc_0701. In this plasmid, the HindIII site 3’ of ColE1 ori is replaced with a BamHI site to facilitate promoter replacement. Colonies were screened by PCR with primers CV615 and CV616 and positive plasmids were verified by sequencing. The lacZ gene was amplified from A. succinogenes 130Z genomic DNA using primers CV624 and CV625, and cloned into pLGZ920* -pAsuc_0701’s XbaI and SacI sites by Gibson cloning. Colonies were screened by PCR using primers CV636 and CV640. A plasmid with the correct lacZ sequence (sequencing with primers listed in Table S2.1) was called pLGZ920*-pAsuc_0701-lacZ. Promoters pAsuc_2109, pAsuc_0391, and pAsuc_0289 were amplified using primers CV 569 and CV570, CV571 and CV572, and CV578 and CV579, respectively (Table S2.1), cut with BamHI and XbaI, and cloned into pLGZ920*-pAsuc_0701-lacZ’s BamHI and XbaI sites (these restrictions excise pAsuc_0701) by ligation using T4 DNA ligase (New England Biolabs, Ipswich, MA). Colonies were screened using CV615 and the reverse primers used for amplifying each promoter fragment. Positive clones were verified by sequencing. All four plasmid constructs were then transformed into strain ΔlacZ. 2.3.11 Construction of strains expressing putative succinate transporters. Asuc_0142, Asuc_1990-1991, Asuc_1999, and Asuc_2058 were amplified from A. succinogenes genomic DNA using primers CV579 and CV580, CV582 and CV586, CV587 and 56 CV588, and CV589 and CV590, respectively. The four PCR fragments were cloned in pLGZ920-ppckA-92-lacZ and pLGZ920-ppckA-103-lacZ cut with XbaI and SacI (which remove lacZ) by Gibson cloning. Colonies were screened by PCR with GZ1302 and GZ1303 and positive clones were verified by sequencing. All constructs were then transformed into A. succinogenes 130Z. 2.3.12 High-performance liquid chromatography (HPLC) analysis of fermentation media from strains expressing transporters. A. succinogenes growth in liquid cultures was monitored by optical density at 660 nm (OD660) on a DU 650 spectrophotometer (Beckman, Fullerton, CA). OD660 values were used to calculate carbon balances. Glucose and fermentation products were quantified in culture supernatants by HPLC (Waters, Milford, MA) using an Aminex HPX-87H column (Bio-rad). Samples were run at room temperature with 4 mM H2SO4 as the eluent at a 0.6 mL min-1 flow rate. Glucose and ethanol were quantified on a Waters 410 differential refractometer. Organic acids were quantified on a Waters 2487 UV detector at 210 nm. 2.4 Results and Discussion 2.4.1 Potential C4-dicarboxylate exporters in A. succinogenes Succinate as a fermentation product is produced by fumarate reductase, whose active site faces the cytoplasm. Succinate has then to be exported out of the cell. A. succinogenes succinate exporters are completely unknown. Table 2.2 lists the putative C4-dicarboxylate transporters identified in the A. succinogenes 130Z genome by BlastP using known C4-dicarboxylate transporters from other bacterial species. A. succinogenes contains homologs of E. coli DcuA, 57 DcuB, DcuC, DauA, CitT, and YjjPB, as well as one homolog of C. glutamicum SucE, and multiple homologs of C. glutamicum DcsT. It also contains a protein of the telluriteresistance/carboxylate transport (TDT) family, up to nine tripartite ATP-independent periplasmic (TRAP family) transporters, and up to five divalent anion:Na+ symporters (DASS family). Table 2.2 A. succinogenes putative C4-dicarboxylate transporters Asuc locus 0020 0023 0074 0142 0146 0147 0148 0156 0157 0158 0183 0270 0271 0272 0273 0304 0366 0367 0368 0715 0716 1063 1163 1164 1165 1482 1568 1577 1578 1579 1781 1921 1922 1923 1957 1958 1988 1989 1990 1991 1999 2058 Putative function Anion transporter, ArsB/NhaD family Putative transporter, aspartate:alanine exchanger family, 32% identical to C. glutamicum SucE C4-dicarboxylate transporter/malic acid transport protein Anaerobic C4-dicarboxylate membrane transporter, 43% identical to E. coli DcuA Possible TRAP C4-dicarboxylate transporter solute receptor, DctP Possible TRAP C4-dicarboxylate transporter, small permease component, DctQ Possible TRAP C4-dicarboxylate transporter, large permease component, DctM Possible TRAP C4-dicarboxylate transporter, large permease component, DctM Conserved hypothetical protein (TRAP, small permease component, DctQ) Possible TRAP C4-dicarboxylate transporter solute receptor, DctP Anion transporter; Anion transporter, ArsB/NhaD family; 33% identical to E. coli CitT Conserved hypothetical protein (TRAP, small permease component, DctQ) Possible TRAP C4-dicarboxylate transporter, large permease component, DctM Possible TRAP C4-dicarboxylate transporter solute receptor, DctP Possible TRAP C4-dicarboxylate transporter solute receptor, DctP Anion transporter, DASS family, involved in fumarate uptake Possible TRAP C4-dicarboxylate transporter solute receptor, DctP Conserved hypothetical protein (TRAP, small permease component, DctQ) Possible TRAP C4-dicarboxylate transporter, large permease component, DctM Putative Thr/Ser exporter, 52% identical to E. coli YjjB Putative Thr/Ser exporter, 50% identical to E. coli YjjP Anaerobic C4-dicarboxylate antiporter, DcuC family Possible TRAP C4-dicarboxylate transporter solute receptor, DctP Possible TRAP C4-dicarboxylate transporter, small permease component, DctQ Possible TRAP C4-dicarboxylate transporter, large permease component, DctM Anion transporter, ArsB/NhaD family; 31% identical to E. coli CitT Anion permease ArsB/NhaD, 36% identical to C. glutamicum DcsT Possible TRAP C4-dicarboxylate transporter, large permease component, DctM Conserved hypothetical protein (TRAP, small permease component, DctQ) Possible TRAP C4-dicarboxylate transporter solute receptor, DctP Hypothetical protein; form of DctA or dicarboxylate transport protein found in many bacterial families Possible TRAP C4-dicarboxylate transporter, large permease component, DctM Possible TRAP C4-dicarboxylate transporter, small permease component, DctQ Possible TRAP C4-dicarboxylate transporter solute receptor, DctP TRAP transporter, 4TM/12TM fusion protein , DctQM TRAP transporter solute receptor, TAXI family TRAP transporter solute receptor, TAXI family UspA domain protein, involved in stress response TRAP transporter, 4TM/12TM fusion permease protein, DctQM TRAP transporter solute receptor, TAXI family Putative anaerobic C4-dicarboxylate transporter, 75% identical to E. coli DcuB (contains a 100-residue insertion) Sulfate permease (sulP) family, 58% identical to E. coli DauA 58 Some of the TRAP transporters could be involved in C4-dicarboxylate transport, but TRAP transporters are involved in solute uptake and are considered mostly unidirectional (28), thus they are unlikely to be involved in succinate efflux. DASS transporters are also involved in solute uptake, and the A. succinogenes DASS transporter SdcA (Asuc_0304) was shown to restore growth on C4-dicarboxylates in a C4-dicarboxylate-transport-negative E. coli strain (29). Other putative C4-dicarboxylate transporters tested (Asuc_0074, _0142, _1063, _1482, and _1999) did not (29). Because DASS transporters work in the direction of a decreasing Na+ gradient, Asuc_0304 is unlikely to be involved in anion efflux. In the same study, qPCR reactions suggested that none of the genes tested (i.e., Asuc_0074, _0142, _0304, _1063, _1482, and _1999) were induced during anaerobic growth on glucose compared to other growth conditions. Asuc_1063 was upregulated 4 and 6 times during aerobic and anaerobic growth on fumarate, respectively, suggesting that this transporter is involved in C4-dicarboxylate uptake as well (29). 2.4.2 Proteomic analysis of A. succinogenes cytoplasmic membrane-enriched fractions Based solely on the putative C4-dicarboxylate transporters identified in the genome and on the functions of the different transporters already characterized in E. coli and C. glutamicum, it is impossible to predict which A. succinogenes putative transporters are the best candidates for succinate export. For this reason, we performed a proteomics analysis of cytoplasmic membraneenriched fractions from A. succinogenes anaerobically grown on different carbon sources (i.e., glucose, fructose, xylose, and glycerol-nitrate). Note that our cytoplasmic membrane fractions were only enriched in cytoplasmic membrane. They did not consist of not entirely purified cytoplasmic membranes. For this reason the fractions still contained outer membrane, 59 cytoplasmic, and periplasmic proteins. Of the top 250 hits, though, 101 were associated with the cytoplasmic membrane, with 14 associated with the outer membrane, and 103 cytoplasmic proteins, indicating that these fractions could still provide good information on cytoplasmic membrane protein abundance. The putative transporter candidates with the most hits (in NSAF format) are listed in Table 2.3. Asuc_1999, a homolog of E. coli DcuB, had the highest number of hits across all conditions tested, followed by Asuc_1990, Asuc_2058, Asuc_0142, and Asuc_1163. Asuc_1990 is the fused permease component (DctQM) of a TRAP-family transporter. Note that the putative solute receptor likely associated with Asuc_1990, Asuc_1991, was detected with high probability (rank 963) in all growth conditions tested, even though it is a periplasmic protein. Asuc_2058 belongs to the sulfate permease family, pfam00916. All characterized members of this family so far are sulfate uptake transporters (30). Asuc_0142 is a DcuA homolog. Asuc_1163 is the putative solute receptor of a TRAP C4-dicarboxylate transporter. Surprisingly, this periplasmic solute receptor was detected in all growth conditions, with a high ranking, while its putative cytoplasmic membrane partners, Asuc_1164 and Asuc_1165, were not detected in any growth condition tested (Table 2.3). Peptides of other putative C4 dicarboxylate transporters were also detected in our samples, but typically at lower levels and not consistently across all growth conditions. Among those, DcuC homolog Asuc_1063 was detected at very low levels across all growth conditions. Asuc_0304, which had been shown to be involved in succinate uptake (29), was not detected in any growth condition. Neither were Asuc_0023, a C. glutamicum SucE homolog, nor Asuc_0715-16, homologs of E. coli YjjBP (Tables 2.3 and 2.4). 60 Table 2.3 Proteomics and transcriptomics data for succinate transporters in A. succinogenes Glucose anaerobic Ranka Protein putative function 46 125 963 216 272 342 554 672 751 918 1122 1138 1258 1289 1372 ND ND ND ND ND ND Anaerobic C4-dicarboxylate transporter, DcuB homolog TRAP-TAXI transporter, 4TM/12TM fusion protein TRAP transporter solute receptor, TAXI family C4-dicarboxylic acid transporter DauA Anaerobic c4-dicarboxylate transporter, DcuA homolog TRAP C4-dicarboxylate transporter solute receptor, DctP TRAP C4-dicarboxylate transporter solute receptor, DctP C4-dicarboxylate ABC transporter C4-dicarboxylate ABC transporter, DASS family C4-dicarboxylate ABC transporter, DcuC homolog Hypothetical protein Anion permease ArsB/NhaD C4-dicarboxylate ABC transporter permease TRAP family transporter membrane component, DctQ TRAP transporter, 4TM/12TM fusion protein, DctQM Corynebacterium glutamicum SucE homolog Dicarboxylate transporter/tellurite-resistance protein Anion transporter, DASS family TRAP transporter solute receptor, TAXI family Anion transporter, similar to Asuc_1482 Anion permease ArsB/NhaD Asuc_ locus 1999 Proteinb Fructose anaerobic RNAc,d Proteinb RNAc Xylose anaerobic Proteinb RNAc Glycerol nitrate Proteinb RNAc Glycerol microaerobic RNAc 0.164 ± 0.015 2698 ± 310 0.171 ± 0.015 1230 ± 98 0.199 ± 0.022 2980 ± 133 0.116 ± 0.017 919 ± 193 1990 0.058 ± 0.023 104 ± 25 195 ± 35 0.038 ± 7E-04 1991 2058 0142 0.037 ± 0.007 0.023 ± 0.007 0.031 ± 0.04 752 ± 222 0.052 ± 5E-04 1094 ± 333 0.050 ± 0.014 702 ± 54 0.028 ± 0.004 887 ± 483 377 ± 17 0.025 ± 0.001 163 ± 9 0.028 ± 0.002 199 ± 9 0.024 ± 0.001 159 ± 19 407 ± 84 0.026 ± 0.008 387 ± 48 0.054 ± 0.007 1025 ± 76 0.072 ± 0.009 377 ± 78 1163 0.009 ±0.003 12 ± 1 0.001 ± 0.002 6±1 0.016 ± 0.002 14 ± 2 0.035 ± 0.007 19 ± 9 11 ± 4 0366 0.002 ± 7E-04 21 ± 3 0.002 ± 2E-05 5±1 0.004 ± 5E-04 31 ± 1 0.003 ± 0.000 18 ± 5 34 ± 2 1781 1482 1063 0272 0020 1577 0367 0.004 ± 0.004 0.002 ±2E-04 0.002 ± 0.004 56 ± 5 98 ± 9 55 ± 2 93 ± 12 87 ± 55 40 ± 3 13 ± 2 0.005 ± 0.001 0.003 ± 0.002 0.007 ± 0.002 1957 6E-04 ± 0.001 0.003 ± 0.002 0.080 ± 0.007 0.003 ± 0.001 120 ± 2 65 ± 2 0.012 ± 0.003 72 ± 1 164 ± 44 0.009 ± 6E-04 236 ± 13 30 ± 2 0.000 ± 0.000 39 ± 2 16 ± 1 47 ± 2 43 ± 13 136 ± 23 36 ± 3 0.001 ± 0.002 87 ± 0 5±1 0.003 ± 0.004 18 ± 1 0.020 ± 0.03 134 ± 47 0.000 ± 0.000 114 ± 36 0.003 ± 0.001 507 ± 87 0.000 ± 0.000 121 ± 7 0.012 ± 0.002 92 ± 22 115 ± 72 64 ± 9 0.006 ± 0.001 13 ± 3 1407 ± 219 88 ± 20 368 ± 85 210 ± 37 234 ± 113 63 ± 5 140 ± 28 44 ± 8 75 ± 23 115 ± 72 57 ± 19 24 ± 1 9±1 13 ± 2 18 ± 0 60 ± 35 13 ± 3 0023 0074 342 ± 38 101 ± 11 146 ± 37 96 ± 18 245 ± 12 98 ± 3 326 ± 140 84 ± 29 33 ± 3 93 ± 3 0304 1988 0183 1568 236 ± 9 140 ± 26 13 ± 2 29 ± 6 212 ± 17 278 ± 61 11 ± 1 23 ± 3 289 ± 9 189 ± 12 12 ± 1 28 ± 3 266 ± 26 179 ± 63 39 ± 9 103 ± 30 309 ± 23 121 ± 30 9±2 247 ± 96 a Rank is the overall rank of each protein in the entire normalized proteomics data set ND: not detected b Proteomics data are in NSAF format c Transcriptome data are expressed per nt d Glu RNA data are average ± standard deviations from three independent biological replicates. 61 2.4.3 Transcript levels of A. succinogenes putative C4-dicarboxylate transporters As seen in Table 2.3, transcript levels for the putative C4 dicarboxylate transporters do not always agree well with detected protein levels, but this observation is not particularly surprising due to post transcriptional regulation, the typically higher stability of proteins compared to RNA, and the large variability in protein detectability by mass spectrometry. This variability can explain why some transporter genes were clearly transcribed (e.g., Asuc_0023, Asuc_0304, or Asuc_1988) but the corresponding protein was not detected in any growth condition. For each putative transporter listed, though, trends in transcript levels were relatively uniform across growth conditions, and four of the top six transporters (including Asuc_1991) identified by proteomics, had the highest transcript levels in anaerobic glucose cultures (i.e., Asuc_1999, Asuc_2058, Asuc_0142, and Asuc_1991). As observed in our proteomics results, Asuc_1999 has the highest transcript levels of all putative C4-dicarboxylate transporters in all growth conditions tested, even in the presence of nitrate. It also ranked very high relative to all transcript levels in the cell (e.g., rank of 101 on glucose). Succinate production during growth on glycerol-nitrate is much lower than during anaerobic growth on glucose (27). Interestingly, asuc_1999’s transcript levels decreased almost 3-fold and its ranking fell almost 4-fold in glycerol-nitrate compared to anaerobic glucose cultures (Table 2.3). Based on these results, Asuc_1999 is a good candidate as a transporter involved in succinate export. Transcript levels for asuc_1990 and asuc_1991 were relatively steady across growth conditions, with transcript levels for asuc_1991 always higher than those for asuc_1990, suggesting that the two genes are not cotranscribed. The facts that the intergenic region between 62 asuc_1990 and asuc_1991 is 336-nt long and that very few RNA reads mapped to this region support this conclusion. Transcripts levels for asuc_2058 were higher in anaerobic glucose cultures than in any other conditions, but they remained at least 100-fold higher than the lowest transcript levels detected for other genes in any conditions. Transcripts levels for asuc_0142 were among the highest for putative C4 dicarboxylate transporters across all growth conditions, in good agreement with the proteomics results. While the Asuc_1163 protein was consistently detected by proteomics in the growth conditions tested, the corresponding transcript levels were very low across all growth conditions, and transcript levels for the putative associated membrane components Asuc_1164 and Asuc_1165 were extremely low as well. Our combined proteomics and RNAseq data suggest that Asuc_1999, Asuc_2058, Asuc_0142, and Asuc_1990-91 could be involved in succinate export. Of these, only Asuc_1999 and Asuc_0142 are homologs of known succinate exporters, DcuB and DcuA, respectively (Table 2.2). Our proteomics and RNAseq data also seem to exclude Asuc_0304, DcuC homolog Asuc_1063, SucE homolog Asuc_0023, and E. coli YjjBP homologs Asuc_0715-16 as succinate exporters. Quantitative real-time PCR studies by Rhie et al (2014) in A. succinogenes grown anaerobically on glucose did not show any induction of genes asuc_0142, asuc_1063, asuc_1482, and asuc_1999. Our RNAseq results do not agree with these results for Asuc_1999 and Asuc_0142 since we saw high transcript levels for both of these genes. 63 2.4.4 Construction of an expression reporter system. E. coli promoters are typically poorly or not functional at all in Pasteurellaceae species. At the beginning of this study, the only A. succinogenes promoter that had been tested was that of the phosphoenolpyruvate carboxykinase gene, pckA, which is a strong, constitutive promoter. No transcriptomic data were available for A. succinogenes to give us a sense of which promoters were strong or weak and when they were functional. Because recombinant membrane protein overexpression often requires a well-calibrated, weak-to-moderately strong promoter, and because this single available A. succinogenes native promoter did not allow for any tunability of gene expression, we sought to decrease the strength of ppckA by first identifying the pckA transcription start site (TSS) and then constructing truncated versions of the promoter. To test these truncated promoters, though, we needed a reporter expression system. lacZ region in strain 130Z lacZ lacZup lacZdown (4.5 kb) ΔlacZ construct in pCR2.1 lacZdown lacZup (1.5 kb) ΔlacZ::icd construct in pCR2.1 and ΔlacZ::icd strain A A lacZup FRT ppckA icd FRT lacZdown (3.3 kb) ΔlacZ strain lacZup 130Z FRT ΔlacZ lacZdown (1.6 kb) (A) ΔlacZ(pLGZ920::lacZ) (B) Figure 2.1 Construction of the ΔlacZ strain. (A) Schematic of the construction procedure. FRT, flippase recognition target site; A, AscI restriction site. (B) A. succinogenes 130Z, ΔlacZ, and ΔlacZ(pLGZ920::lacZ) strains grown on LB-glucose supplemented with X-gal (40 µg mL-1). 64 To test promoter strength in A. succinogenes, we used lacZ as the reporter gene. A lacZ deletion mutant of A. succinogenes was constructed (strategy described in Figure 2.1A) to be used as the background strain. The ΔlacZ strain was confirmed by testing lacZ expression on LB-glucose plates supplemented with 5-bromo-4-chloro-3-indolyl-β-D-galactoside (X-gal, 40 µg mL-1) (Figure 2.1B). Plasmid pLGZ920::lacZ, with lacZ under control of ppckA, was used as the positive control to test the reporter system (Figure 2.1B). 2.4.5 Identification of ppckA’s TSS by RLM-RACE and construction of truncated ppckA promoters Out of the 14 clones sequenced, six clones indicated the TSS (+1) of pckA as a T nucleotide 32 nt upstream of the start codon. Sequences similar to E. coli promoter -10 and -35 consensus sequences were found 14 nt and 27 nt upstream of the TSS, respectively, with a 21-nt spacer sequence between the probable -10 and -35 sites (Figure 2.2). ATTTGAAACGGATCACAAATCATGAAAAAAATACGTTCAAATTAGAACTAATTATCGAAAATTTGATCTA GTTAACATTTTTTAGGTATAAATAGTTTTAAAATAGATCTAGTTTGGATTTTTAATTTTAAATTATCAAT GAGGTGAAGTATGACTGACTTAAACAAACTCGTTAAAGAACT Figure 2.2 Identification of the pckA transcription start site. 5’ region of the A. succinogenes pckA gene. Yellow highlight: probable -35 region; green highlight: probable -10 region; red and bold: transcription start site; underline: ribosome binding site; bold characters: start of the pckA coding region. In our earlier studies, we routinely used a 231-nt sequence upstream of the pckA start codon to express genes under the control of ppckA. We built truncated versions of this 231-nt sequence to decrease the strength of the promoter (Figure 2.3), and tested the strength of these truncated promoters using lacZ as the reporter gene. Activity of the truncated promoters decreased with the 65 length of the promoter region, and ranged from 1.4-fold (ppckA-164 promoter) to 20-fold lower (ppckA-92 promoter) than ppckA (Figure 2.4). Almost no β-galactosidase activity was detected with promoter ppckA-75 (data not shown), which suggests that the identified putative -35 region is indeed needed for promoter function. Figure 2.3 Reporter constructs with lacZ under control of truncated versions of pckA’s 5’UTR. H, HindIII; X, XbaI; yellow highlight: probable -35 region; green highlight: probable 10 region; red and bold: transcription start site; underline: ribosome binding site; italics: indicate restriction sites—HindIII (5’ end) and XbaI (3’ end); vertical arrows indicate start of truncated promoters. Drawings of DNA fragments are not to scale. 66 2.4.6 Identification of promoters of different strengths from the transcriptomics results. To expand the expression range of our promoter library toward expression levels lower than from ppckA-92, we used our RNAseq results. Candidate promoters were selected from the genes whose transcript levels did not vary between growth conditions using the RNAseq results from anaerobic cultures grown on glucose, xylose, fructose, and mannose, as well as glycerol microaerobic cultures, which all favor succinate production. Genes whose log2 fold-change was between -0.2 and 0.2 in all growth conditions compared to glucose were compiled and their transcript levels (using expression levels per nucleotide) (Table S2.2) were ranked, together with that of pckA, for further analysis. Candidates (individual genes or the first genes in an operon) were selected that covered a large range of decreasing strengths compared to ppckA. Promoters of candidate genes asuc_0391, asuc_0289, asuc_0701, and asuc_2109 were selected, whose transcript levels were 16-, 24-, 29-, and 62-fold lower than pckA’s. DNA fragments 247 bp upstream of candidate genes asuc_0701, asuc_0289, and asuc_0391 and 261-bp upstream of asuc_2109 were cloned in front of lacZ to be tested as promoters. Promoter strength was tested by β-galactosidase assays using ΔlacZ as the expression strain, with lacZ under control of ppckA as the positive control (Figure 4). Promoter pAsuc_0701 was the weakest promoter, 209-fold weaker than ppckA, which was the strongest. To the best of our knowledge this is the first promoter library constructed for A. succinogenes 130Z, making it easier to test expression of different genes in this strain for succinate production. 67 Specific activity (µmol/min/mg) 100000 90000 80000 70000 60000 50000 40000 30000 20000 10000 8000 7000 6000 5000 4000 3000 2000 1000 0 1 2 3 4 5 6 7 8 9 10 Figure 2.4 β-galactosidase activity of promoter reporter constructs. All promoter constructs were tested in the ΔlacZ strain. 1, ΔlacZ(pLGZ920); 2, ΔlacZ(pLGZ920*-pAsuc_0701-lacZ); 3, ΔlacZ(pLGZ920*-pAsuc_2109-lacZ); 4, ΔlacZ(pLGZ920*-pAsuc_0391-lacZ); 5, ΔlacZ(pLGZ920-ppckA92-lacZ); 6, ΔlacZ(pLGZ920*-pAsuc_0289-lacZ); 7, ΔlacZ(pLGZ920-ppckA-103-lacZ); 8, ΔlacZ(pLGZ920-ppckA-134-lacZ); 9, ΔlacZ(pLGZ920*-ppckA-164-lacZ); 10, ΔlacZ(pLGZ920-lacZ). Results are the average activity ± standard deviation from three independent biological replicates. 2.4.7 Overexpression of putative succinate transporter candidates in A. succinogenes and succinate production. Succinate transporter candidates with the most hits in the proteomics analysis and the highest transcript levels were over-expressed to determine their effects on succinate production. We focused on Asuc_1999, Asuc_1990-1991, Asuc_2058, and Asuc_0142. Even though 68 Asuc_1990 and Asuc_1991 seemed to be transcribed independently, they are likely part of the same transport system, so the two genes were expressed in a single construct. For each putative transporter, we tested several weak promoters to maximize protein expression while avoiding toxicity to the cells. Expression of all four transporters under control of ppckA-134 was toxic to the cells, thus we tested expression of these proteins under control of ppckA-92 and ppckA-103. Asuc_2058 and Asuc_1990-1991 expressed under control of ppckA-103 proved toxic to the cells as well, but expression of Asuc_0142 and Asuc_1999 under the same promoter did not inhibit cell growth. Promoter ppckA-92 allowed expression of all four transporters without inhibiting growth of the 130Z recombinant strain. Fermentation balances for A. succinogenes 130Z carrying the six expression constructs are shown in Table 2.4. All six expression constructs caused A. succinogenes 130Z’s succinate yields to increase significantly, compared to the empty vector, in anaerobic glucose cultures. Much of the succinate yield increase seemed to be at the expense of biomass production. Note that expressing Asuc_1999 and Asuc_0142 genes under control of ppckA-103 increased the succinate yield compared to ppckA-92 for both transporters, although the increase was not statistically significant. The strain with the highest succinate yield, 130Z (pLGZ920-ppckA-103-Asuc_1999), also produced almost the highest acetate yield, but produced the least biomass. In contrast, the second best succinate producing strain, 130Z (pLGZ920-ppckA-92-Asuc_1990-1991), produced almost as much succinate, plus it did not produce more acetate than the control strain, making 130Z (pLGZ920-ppckA-92-Asuc_1990-1991) possibly a more suitable candidate for succinate production and further engineering. 69 Table 2.4 Fermentation balances of strain 130Z over-expressing candidate succinate transporters. Products (mmol/100 mmol glucose consumed) Strain Succinate Formate Acetate Ethanol 130Z (pLGZ920) 130Z (pLGZ920-ppckA-92-Asuc_1999) 130Z (pLGZ920-ppckA-92-Asuc_0142) 130Z (pLGZ920-ppckA-103-Asuc_1999) 130Z (pLGZ920-ppckA-103-Asuc_0142) 130Z (pLGZ920-ppckA-92-Asuc_2058) 130Z (pLGZ920-ppckA-92-Asuc_1990-91) 47.3 ± 2.3 55.0 ± 0.8 53.7 ± 1.4 58.5 ± 2.8 56.4 ± 1.1 55.7 ± 2.4 57.2 ± 2.2 98.0 ± 5.4 103 ± 1 98.5 ± 2.4 101 ± 6 101 ± 3 98.1 ± 5.2 96.9 ± 0.6 65.5 ± 2.8 72.4 ± 1.1 64.6 ± 2.1 73.0 ± 3.3 73.9 ± 1.6 68.4 ± 2.9 65.0 ± 0.8 30.6 ± 2.8 28.5 ± 2.0 25.5 ± 4.2 25.2 ± 3.2 29.8 ± 0.9 28.5 ± 0.9 26.8 ± 1.7 CO2 Biomassa 0.0 ± 0.0 193 ± 10 0.0 ± 0.0 157 ± 2 0.0 ± 0.0 188 ± 6 0. 0± 0.0 138 ± 5 0.5 ± 0.3 152 ± 8 0.0 ± 0.0 142 ± 2 0.0 ± 0.0 154 ± 2 % increased Carbon succinate recoveryb yield 104 ± 5 104 ± 1 104 ± 3 101 ± 4 105 ± 3 100 ± 4 101 ± 1 NA 16** 14* 24** 19** 18* 21** Doubling time 2.04 ± 0.03 2.22 ± 0.05 2.27 ± 0.02 2.74 ± 0.2 2.46 ± 0.01 2.49 ± 0.06 2.26 ± 0.05 Results are an average of three biological replicates ± standard deviations. a Biomass was determined using assumed values of 567 mg dry cell weight/mL per OD660(31) and a cell composition of CH2O0.5N0.2 (24.967 g/mol) (32) b Carbon balance is the carbon in products/carbon in glucose consumed. It is assumed that one CO2 is fixed for each molecule of succinate produced. CO2 was calculated using the following formula: CO2 (in mM) = acetate (in mM) + ethanol (in mM) - formate (in mM) * Significantly different from 130Z (pLGZ920) (p < 0.05, two-tailed student’s t-test) ** Significantly different from 130Z (pLGZ920) (p < 0.01, two-tailed student’s t –test) 70 When attempting to express all the constructs in the ΔpflB strain (a pyruvate-formate lyase deletion mutant) (21), which does not produce any formate, we were either unable to obtain any transformants or the strains grew extremely poorly, indicating that these constructs are toxic in ΔpflB and that weaker promoters are likely needed in this strain (data not shown). 2.4.8 The A. succinogenes succinate efflux transporters differ from those of E. coli and C. glutamicum A. succinogenes contains homologs of many E. coli succinate transporters—DcuA, DcuB, DcuC, DauA, CitT, and YjjPB. It also contains one homolog of C. glutamicum SucE, along with multiple homologs of C. glutamicum DcsT. E. coli DcuC can perform different functions, but, in particular, it has been shown to participate in succinate efflux under anaerobic conditions (9). Asuc_1063 is the A. succinogenes DcuC homolog, but it does not show high transcriptional levels or proteomics hits in our studies, eliminating it as a succinate efflux transporter. E. coli YjjP and YjjB constitute a major succinate efflux transporter (17) but Asuc_0715 and Asuc_0716, which are 52% and 50% identical to YjjB and YjjP, respectively, were not detected in our proteomics analysis. Golby et al. (33) determined that E. coli DcuB is subject to catabolite repression, repressed by nitrate, and strongly induced by C4-dicarboxylates. These authors suggested that it is mostly involved in C4-dicarboxylate transport during fumarate respiration. A quintuple or sextuple deletion mutant, including ΔdcuB, did not show decreased succinate efflux (13), but then a single dcuB mutant affected succinate efflux during growth on glucose (9), thus DcuB could contribute to succinate efflux in E. coli. Asuc_1999, a putative anaerobic C4-dicarboxylate transporter that is 75% identical to E. coli DcuB increases the succinate yield in A. succinogenes when 71 overexpressed. Asuc_1999 had the most hits in our proteomics analysis and had the highest transcript levels of all putative C4-dicarboxylate transporters across all conditions tested, suggesting that it plays a major role in succinate efflux in A. succinogenes. Overexpressing Asuc_1990-1991 in A. succinogenes caused an increase in succinate yield. This result is rather surprising, since Asuc_1990-1991 belongs to the TRAP family of transporters, which are primarily involved in uptake transport. Asuc_1990-1991 had the second highest hits in our proteomics analysis and had high transcriptional levels across all the conditions we tested. Asuc_0142, an anaerobic C4-dicarboxylate membrane transporter, is 43% identical to E. coli DcuA. Overexpression of Asuc_0142 in A. succinogenes also caused an increase in succinate yield. E. coli DcuA is known to function as both a succinate uptake and efflux transporter, as confirmed using double mutants expressing only one of the DcuA, DcuB, or DcuC carriers at a time (8). Asuc_2058, showing similarity to E. coli DauA, also led to an increase in succinate yield upon overexpression. This result suggests that Asuc_2058 has a function completely different from that of E. coli DauA, which has been shown to be the main succinate uptake transporter under acidic conditions and to be inactive at pH 7 (10). SucE has been identified as one of the C. glutamicum transporters involved in succinate efflux under microaerobic and anaerobic conditions. Asuc_0023, a homolog of C. glutamicum SucE had moderate transcript levels but was not detected in our proteomics studies, indicating that the C. glutamicum and A. succinogenes homologs do not have the same function. 72 Thus succinate efflux transporters in A. succinogenes are different from the major succinate exporters of E. coli and C. glutamicum, and even include transporters (e.g., TRAP-family transporters Asuc_1990-91) that have so far never been shown to function in export. 2.5 Conclusion This study is the first to focus on A. succinogenes dicarboxylate efflux transporters with the aim of increasing succinate production. Combining proteomics and transcriptomics approaches, we identified the four putative C4-dicarboxylate transporters with the highest expression levels. Overexpressing all four transporters individually increased succinate production. While, based on our expression data, Asuc_1999 is a likely candidate to be a major succinate exporter, we cannot identify a single major succinate exporter in A. succinogenes. Our results suggest instead that more than one transporter is involved in succinate export in A. succinogenes, similar to the situation observed in E. coli. Knockout mutants of these transporters (construction unsuccessful so far) would shed more light on the relative involvement of these transporters in succinate efflux in A. succinogenes. We also developed a library of A. succinogenes promoters covering a large expression range across multiple growth conditions, which can be used for modulating protein expression in A. succinogenes as needed for strain engineering. 2.6 Acknowledgements I thank Dr. Claire Vieille and Peggy Wolf for carrying out all work related to proteomics. I thank Dr. Nikolas McPherson for obtaining and analyzing all RNA sequencing data. 73 APPENDIX 74 Table A2.1 Primers used in this study Primer P1 P2 CV376 CV377 CV443 CV469 CV470 CV471 CV473 CV474 CV475 CV476 CV569 CV570 CV571 CV572 CV577 CV578 CV579 CV580 CV586 CV582 CV587 CV588 CV589 CV590 CV615 CV616 CV617 CV618 CV624 CV625 GZ1302 GZ1303 CV431 Sequencea (restriction site) CGTGGTTAACGTCCCTTTATCG GGTTTCTTCCTCGAAAAGTTGTTC CTTGCCAAACCGACCGAAAG TATTGATAATGAAAATCCGACCGCACTTGGCAGTACCGGCGTATTCCTC b CACACCATTCCCAAACAAAAC GACCATGATTACGCCAAGCTTTCTGCTAATCCTGTTACCAGTGG TAGGTTGGCAGAATCATCTAGATCACCTCATTGATAATTTAAAATTAAAAATCCAAACT AGATCTATTTTAAAAC AAGCTTTCTTCCGCTTCCTCGCTCACTG GAGGAAGCGGAAGAAAGCTTCGAAAATTTGATCTAGTTAACATTTTTTAGG GAGGAAGCGGAAGAAAGCTTGAACTAATTATCGAAAATTTGATCTAG GAGGAAGCGGAAGAAAGCTTCACAAATCATGAAAAAAATACGTTC GAGGAAGCGGAAGAAAGCTTCATTTACCGCCATAAAAATTTGAAAC TGAGCGAGGAAGCGGAAGAGGATCCGGTCAAACTCCTACGAATTTC AAGTAGGTTGGCAGAATCATCTAGAGCTCACCAACAGGCTTGA TGAGCGAGGAAGCGGAAGAGGATCCGAGAGAACAGGACAACAGTTTTATTG AAGTAGGTTGGCAGAATCATCTAGAAATAATAACTCTTAATATAGAAAAAAACGATTG TGAGCGAGGAAGCGGAAGAGGATCCGACGAGCTACTCCCTGGTTTG AAGTAGGTTGGCAGAATCATCTAGAGATCCTTTTAAAAAAAACGATGAAAAAAT TAAATTATCAATGAGGTGATCTAGATGACTGCAATGTTTATTATCC ACGGCCAGTGAATTCGAGCTCTTAGATAAACAGATTCGCAAATAAC TAAATTATCAATGAGGTGATCTAGATGAAAAAATTATTTAAACTTTCTCTTGTC ACGGCCAGTGAATTCGAGCTCTATTCCGGACGACGACG TAAATTATCAATGAGGTGATCTAGATGGATTTTTTGATGAATCTAAG ACGGCCAGTGAATTCGAGCTCTTAAAGATAACCGTATAAGCCTG TAAATTATCAATGAGGTGATCTAGATGCTAAATAAATGGTTTTTAACC ACGGCCAGTGAATTCGAGCTCTTAATGACGTAATTCCGATTC GACCATGATTACGCCAAGCTTTCTGCTAATCCTGTTACCAGTGGC ATCACCTTAGGATCCCCGCATCAGGCGCTCTTC TGATGCGGGGATCCTAAGGTGATTTATAGTCTGGACGG GTACCCGGGGATCCTCTAGATTCACCTCGAAACAGATAAAAAAATC AAATTATCAATGAGGTGATCTAGATGATTCTGCCAACCTACTTTGAAAATCC CCAGTGAATTCGAGCTCTTATTCAAAGTGAATATCGAAACGACAGTCAAATTTTG CGTTGTAAAACGACGGCC AATTTTAAATTATCAATGAGGTG GGTTTCGCCCACTCGTATTCC 75 Specificityb (directionc) 5’RLM-RACE pckA-specific outer primer (R) 5’RLM-RACE pckA-specific inner primer (R) ΔlacZ-fusion, nested (F) ΔlacZ-fusion, nested (R) Truncated ppckA constructs (R) colE1 ori (F) ppckA with lacZ overhangs (R) ColE1 ori (R) ppckA-92 (F) ppckA-103 (F) ppckA-134 (F) ppckA-164 (F) pAsuc_2109 (F) pAsuc_2109 (R) pAsuc_0391 (F) pAsuc_0391 (R) pAsuc_0289 (F) pAsuc_0289 (R) Asuc_0142 (F) Asuc_0142 (R) Asuc_1990-1991 (F) Asuc_1990-1991 (R) Asuc_1999 (F) Asuc_1999 (R) Asuc_2058 (F) Asuc_2058 (R) ColE1 ori (F) ColE1 ori (R), replaces ori’s 3’ HindIII site with BamHI pAsuc_0701 (F) pAsuc_0701 (R) lacZ Gibson cloning (F) lacZ Gibson cloning (R) pLGZ901 specific primer downstream of SacI site (R) pLG901 specific primer upstream of XbaI site (F) lacZ sequencing primer-1 Table A2.1 (cont’d) CV445 GCTCATATGAAATTACATAGCCTGTTTTCGGATCGC CV446 CTATGTAATTTCATATGAGCAGGCGTTAGTTGATTTG CV636 GGATATGGTGAAAACTTCGAAG CV637 CCATTGTTTGGTTTACTCTCC a F, forward primer; R, reverse primer b Primer P4 contains the USS ACCGCACTT lacZ sequencing primer-2 lacZ sequencing primer-3 lacZ sequencing primer-4 lacZ sequencing primer-5 76 Table A2.2 Transcript levels of candidate genes with low variation across several conditions, compared to pckA Gene ID Average expression Expression Fold lower expression per nucleotide compared to pckA than pckA Asuc_2109 Asuc_0708 Asuc_1818 Asuc_1044 Asuc_1536 88 93 101 125 130 0.016 0.017 0.019 0.023 0.024 61.7 58.7 53.8 43.4 41.8 Asuc_1879 Asuc_1430 Asuc_0701 145 170 189 0.027 0.031 0.035 37.5 32.1 28.8 Asuc_0289 Asuc_1432 Asuc_2046 Asuc_1945 Asuc_1699 Asuc_1071 227 228 246 272 285 310 0.042 0.042 0.045 0.050 0.052 0.057 24.0 23.9 22.2 20.0 19.1 17.6 Asuc_0391 Asuc_1037 Asuc_0983 Asuc_1877 Asuc_2007 Asuc_2066 332 469 472 483 549 1423 0.061 0.086 0.087 0.089 0.101 0.261 16.4 11.6 11.5 11.3 9.9 3.8 77 REFERENCES 78 REFERENCES 1. Zeikus JG, Jain MK, Elankovan P. 1999. Biotechnology of succinic acid production and markets for derived industrial products. Appl Biochem Biotechnol 51:545-552. 2. Sanchez AM, Bennett GN, San KY. 2005. Efficient succinic acid production from glucose through overexpression of pyruvate carboxylase in an Escherichia coli alcohol dehydrogenase and lactate dehydrogenase mutant. Biotechnol Prog 21:358-365. 3. Jantama K, Haupt MJ, Svoronos SA, Zhang XL, Moore JC, Shanmugam KT, Ingram LO. 2008. Combining metabolic engineering and metabolic evolution to develop nonrecombinant strains of Escherichia coli C that produce succinate and malate. Biotechnol Bioeng 99:1140-1153. 4. Litsanov B, Brocker M, Bott M. 2012. Toward homosuccinate fermentation: metabolic engineering of Corynebacterium glutamicum for anaerobic production of succinate from glucose and formate. Appl Environ Microbiol 78:3325-3337. 5. Ito Y, Hirasawa T, Shimizu H. 2014. Metabolic engineering of Saccharomyces cerevisiae to improve succinic acid production based on metabolic profiling. Biosci Biotechnol Biochem 78:151-159. 6. Six S, Andrews SC, Unden G, Guest JR. 1994. Escherichia coli possesses two homologous anaerobic C4-dicarboxylate membrane transporters (DcuA and DcuB) distinct from the aerobic dicarboxylate transport system (Dct). J Bacteriol 176:64706478. 7. Engel P, Kramer R, Unden G. 1994. Transport of C4-dicarboxylates by anaerobically grown Escherichia coli. Energetics and mechanism of exchange, uptake and efflux. Eur J Biochem 222:605-614. 8. Zientz E, Six S, Unden G. 1996. Identification of a third secondary carrier (DcuC) for anaerobic C4-dicarboxylate transport in Escherichia coli: roles of the three Dcu carriers in uptake and exchange. J Bacteriol 178:7241-7247. 9. Janausch IG, Unden G. 1999. The dcuD (former yhcL) gene product of Escherichia coli as a member of the DcuC family of C4-dicarboxylate carriers: lack of evident expression. Arch Microbiol 172:219-226. 10. Karinou E, Compton EL, Morel M, Javelle A. 2013. The Escherichia coli SLC26 homologue YchM (DauA) is a C4-dicarboxylic acid transporter. Mol Microbiol 87:623640. 79 11. Pos KM, Dimroth P, Bott M. 1998. The Escherichia coli citrate carrier CitT: a member of a novel eubacterial transporter family related to the 2-oxoglutarate/malate translocator from spinach chloroplasts. J Bacteriol 180:4160-4165. 12. Davies SJ, Golby P, Omrani D, Broad SA, Harrington VL, Guest JR, Kelly DJ, Andrews SC. 1999. Inactivation and regulation of the aerobic C4-dicarboxylate transport (dctA) gene of Escherichia coli. J Bacteriol 181:5624-5635. 13. Janausch IG, Kim OB, Unden G. 2001. DctA- and Dcu-independent transport of succinate in Escherichia coli: contribution of diffusion and of alternative carriers. Arch Microbiol 176:224-230. 14. Chen J, Zhu X, Tan Z, Xu H, Tang J, Xiao D, Zhang X. 2014. Activating C4dicarboxylate transporters DcuB and DcuC for improving succinate production. Appl Microbiol Biotechnol 98:2197-2205. 15. Huhn S, Jolkver E, Kramer R, Marin K. 2011. Identification of the membrane protein SucE and its role in succinate transport in Corynebacterium glutamicum. Appl Microbiol Biotechnol 89:327-335. 16. Fukui K, Koseki C, Yamamoto Y, Nakamura J, Sasahara A, Yuji R, Hashiguchi K, Usuda Y, Matsui K, Kojima H, Abe K. 2011. Identification of succinate exporter in Corynebacterium glutamicum and its physiological roles under anaerobic conditions. J Biotechnol 154:25-34. 17. Fukui K, Nanatani K, Hara Y, Yamakami S, Yahagi D, Chinen A, Tokura M, Abe K. 2017. Escherichia coli yjjPB genes encode a succinate transporter important for succinate production. Biosci Biotechnol Biochem 81:1837-1844. 18. Kim P, Laivenieks M, McKinlay J, Vieille C, Zeikus JG. 2004. Construction of a shuttle vector for the overexpression of recombinant proteins in Actinobacillus succinogenes. Plasmid 51:108–115. 19. Ausubel FM, Brent R, Kingston RE, Moore DD, Seidman JG, Smith JA, Struhl K. 1993. Current Protocols in Molecular Biology. Greene Publishing & Wiley-Interscience, New York. 20. McKinlay JB, Zeikus JG, Vieille C. 2005. Insights into Actinobacillus succinogenes fermentative metabolism in a chemically defined growth medium. Appl Environ Microbiol 71:6651-6656. 21. Joshi RJ, Schindler BD, McPherson NR, Tiwari K, Vieille C. 2014. Development of a markerless knockout method for Actinobacillus succinogenes. Appl Environ Microbiol 80:3053-3061. 80 22. Lee SJ, Song H, Lee SY. 2006. Genome-based metabolic engineering of Mannheimia succiniciproducens for succinic acid production. Appl Environ Microbiol 72:1939–1948. 23. Tai SP, Kaplan S. 1985. Intracellular localization of phospholipid transfer activity in Rhodopseudomonas sphaeroides and a possible role in membrane biogenesis. J Bacteriol 164:181-186. 24. Zhang Y, Wen Z, Washburn MP, Florens L. 2010. Refinements to label free proteome quantitation: how to deal with peptides shared by multiple proteins. Anal Chem 82:22722281. 25. McPherson N. 2017. Novel Insights Into Sugar and Succinate Metabolism of Actinobacillus succinogenes from Strains Evolved for Improved Growth on Lignocellulose Hydrolysate Sugars. Doctoral. Michigan State University. 26. Johnson BK, Scholz MB, Teal TK, Abramovitch RB. 2016. SPARTA: Simple program for automated reference-based bacterial RNA-seq transcriptome analysis. BMC Bioinformatics 17:66. 27. Schindler BD, Joshi RJ, Vieille C. 2014. Respiratory glycerol metabolism of Actinobacillus succinogenes 130Z for succinate production. J Ind Microbiol Biotechnol 41:1339-1352. 28. Mulligan C, Fischer M, Thomas GH. 2011. Tripartite ATP-independent periplasmic (TRAP) transporters in bacteria and archaea. FEMS Microbiol Rev 35:68-86. 29. Rhie MN, Yoon HE, Oh HY, Zedler S, Unden G, Kim OB. 2014. A Na+-coupled C4dicarboxylate transporter (Asuc_0304) and aerobic growth of Actinobacillus succinogenes on C4-dicarboxylates. Microbiology 160:1533-1544. 30. Aguilar-Barajas E, Diaz-Perez C, Ramirez-Diaz MI, Riveros-Rosas H, Cervantes C. 2011. Bacterial transport of sulfate, molybdate, and related oxyanions. Biometals 24:687707. 31. McKinlay JB, Shachar-Hill Y, Zeikus JG, Vieille C. 2007. Determining Actinobacillus succinogenes metabolic pathways and fluxes by NMR and GC-MS analyses of 13Clabeled metabolic product isotopomers. Metab Eng 9:177–192. 32. van der Werf MJ, Guettler MV, Jain MK, Zeikus JG. 1997. Environmental and physiological factors affecting the succinate product ratio during carbohydrate fermentation by Actinobacillus sp. 130Z. Arch Microbiol 167:332–342. 33. Golby P, Kelly DJ, Guest JR, Andrews SC. 1998. Transcriptional regulation and organization of the dcuA and dcuB genes, encoding homologous anaerobic C4dicarboxylate transporters in Escherichia coli. J Bacteriol 180:6586-6596. 81 Chapter 3 Identification of sRNA in Actinobacillus succinogenes 130Z 82 3.1 Abstract Small RNAs (sRNAs) are powerful tools in metabolic engineering and synthetic biology. Synthetic sRNAs are being designed using Hfq-dependent sRNA scaffolds. Hfq is a chaperone protein known to facilitate binding of sRNAs to their mRNA targets and also protecting them from degradation. Recent studies have focused on the identification of sRNAs in a few commercially-relevant organisms, particularly Escherichia coli, to develop them as tools for metabolic engineering. Actinobacillus succinogenes is one such industrially-relevant organism, known to be one the best natural succinate producers. In this study, we identified sRNAs in A. succinogenes to gain better insight into their sequence, size, structure, and function. Our goal was to gain enough information to be able to design synthetic sRNAs that would act as posttranscriptional regulators in our strain. We performed RNAseq analysis of the wild-type strain grown microaerobically on glycerol and anaerobically on glucose. RNAseq data were analyzed using Rockhopper and manually using the Integrated Genomics Viewer. Using ARNold, we identified sRNAs with Rho-independent terminators, a key characteristic of Hfq-dependent sRNAs. We also looked for other known features of Hfq-dependent sRNAs, which allowed us to narrow our focus down to only a few candidates for our synthetic sRNA design. Scaffolds from two sRNAs—smRNA8 and smRNA28, were used in synthetic sRNA constructs to test inhibition of β-galactosidase expression in A. succinogenes. A 32% decrease in β-galactosidase activity was seen with one of the constructs. The target binding region for lacZ was then replaced by the target binding regions for ackA and pta. One of four constructs we tested caused a 14% decrease in acetate yield. To our knowledge, this is the first study identifying sRNAs in A. succinogenes. We have also provided a proof of concept for using synthetic sRNAs in A. succinogenes as a 83 metabolic engineering tool. However, more optimization of these synthetic sRNAs in A. succinogenes is needed to develop a robust and tunable system for future engineering efforts. 3.1 Introduction Small RNAs (~30-500 nucleotides) are regulatory RNAs known to regulate mRNA transcript levels and translation in bacteria. Since they were first discovered 30 years ago, a large number of small RNAs (sRNAs) have been identified in several bacterial species, in part due to advances in deep RNA sequencing and computational tools (1, 2). It is now known that sRNAs have a significant role in gene regulation, but understanding of their mechanisms of action lags behind, and many of their targets remain unknown. sRNAs are most commonly classified by their mechanism of action as reviewed by Gottesman et al. (3). Antisense sRNAs can be classified as cis-encoded sRNAs or trans-encoded sRNAs. Most commonly studied are transencoded antisense sRNAs (encoded by a locus different from their target[s]) that bind to their target mRNA(s) with limited and varying degrees of complementarity. Due to their limited binding capabilities these sRNAs may require the assistance of Sm-like protein Hfq (4) Cisencoded antisense sRNAs are usually encoded by the same locus as the mRNA but on the opposite strand, making them entirely complementary to their target mRNA, in contrast to transencoded antisense sRNAs. Riboswitches are another group of cis-encoded and -acting RNA elements, which sense small molecules and change conformation to allow or block translation (5). In some cases they have been shown to act in trans (6). CRISPRs (clustered regularly interspaced short palindromic repeats) are a newly identified group of sRNAs playing an important role in bacterial immunity against foreign DNA (7), (3). Many small RNAs are also synthesized from the 3’ regions of mRNA. These small RNAs are produced by mRNA 84 processing or transcribed separately with a shared terminator (8). A couple of studies have shown that 3’-UTR derived sRNAs acted as trans regulators for mRNAs (9, 10). DapZ was the first example of a small RNA that was transcribed by an independent promoter from the sense 3’UTR of a gene, dapB (10). DapB is responsible for the catalyzing the second step of lysine biosynthesis (11). The function of DapZ was found to be regulation of major ABC transporters. Thus the gene and the small RNA derived from the sense 3’UTR of the gene had completely different functions. Another example; CpxQ is a sRNA that is generated by RNase E cleavage of CpxP, a stress chaperone. CpxQ, an Hfq-dependent sRNA was found to repress production of several inner membrane proteins (9). However, in this case, the gene and sRNA both were transcribed from the same promoter and therefore transcribed under the same conditions. Most of the studies that have identified and characterized sRNAs have focused on bacterial members of the Enterobacteriaceae family, mainly in Escherichia coli. Sixty one sRNAs have been annotated in Ecogene for E. coli, but many more have been predicted. Many research groups are now investigating how these newly identified sRNAs function in gene regulation. A few studies have been conducted on sRNAs in other bacterial families. Most of them focus on human pathogens, such as Clostridium difficile, Neisseria gonorrhoeae, Acinetobacter baumannii, Pseudomonas aeruginosa, Rickettsia prowazekii, Burkholderia cenocepacia, Streptococcus pneumoniae, Listeria monocytogenes, and Mycobacterium tuberculosis, to identify virulence-related sRNAs (12-20). Very few studies have focused on the Pasteurellaceae family, even though many members of this family are known to be veterinary pathogens (e.g., Actinobacillus pleuropneumoniae, Haemophilus parasuis, Mannheimia paralytica, and Pasteurella multocida), and two, Haemophilus influenzae and Aggregatibacter actinomycetemcomitans, cause disease in humans (19). A majority of the Pasteurellaceae species 85 are difficult to culture outside their host environment, and they are known to have relatively small genomes possibly due to adaptation to their host environment (21). Very little is known about the sequence, structure, and function of sRNAs in Pasteurellaceae, apart from a few studies in H. influenzae and A. pleuropneumoniae. HrrF was the first sRNA identified in a Haemophilus species. It is expressed at high levels when iron availability is low, is regulated by Fur, and is conserved in several other Pasteurellaceae species (22). Eighteen sRNAs have been identified in H. influenzae by RNA sequencing (RNAseq), of which HrrF has been experimentally verified (22). Of these eighteen sRNAs, seven belong to known sRNA families in the Rfam database (23). Seven sRNAs possibly interacting with the carbon storage regulatory protein CsrA were computationally identified in Haemophilus spp. genomes (24). Three sRNAs were also identified in A. actinomycetemcomitans that are regulated by iron and Fur. Most recently, A. pleuropneumoniae sRNAs were identified using four prediction algorithms—BLAST/Rfam, SIPHT, Infernal, and RNAz. sRNAs predicted by two or more programs were considered to be candidates for further analysis. Seventeen of the 23 sRNAs identified using this approach were confirmed by northern blotting (25). Actinobacillus succinogenes is a non-pathogenic member of the Pasteurellaceae family. This capnophilic organism, isolated from a cow’s rumen, is among the best natural succinate producers (26). We are interested in engineering A. succinogenes for succinate production. Although knockout methods and overexpression vectors exist, they are not always the best means to engineer the organism for industrial level production. Many a times a more fine-tuned approach is needed. Synthetic sRNAs have being designed and used to modulate gene expression in a couple of industrially relevant microorganisms (27, 28). We sought to identify sRNAs in A. 86 succinogenes and gain further information on their sequences, structure, size, and function, to then be able to design synthetic sRNAs for engineering purposes. In this study, we carried out deep sequencing of sRNAs from A. succinogenes grown on two different carbon sources—glucose and glycerol. We analyzed the RNAseq data using Rockhopper and by manually reviewing the data on the Integrative Genomics Viewer (IGV). We compared the data obtained from our sequencing results to data obtained from the predictions of a few computational programs. We also sought to identify Hfq-dependent sRNAs in A. succinogenes, and tested them as scaffolds in the design of synthetic regulatory sRNAs. 3.2 Materials and Methods 3.2.1 Bacterial strains and growth conditions. E. coli strains (Table 3.1) were cultivated in lysogeny broth (LB) and plates and were supplemented with 100 µg mL-1 ampicillin for plasmid maintenance where required. A. succinogenes type strain 130Z (ATCC 55618) was grown in 60-mL AM3 medium containing 50 mM glucose and 150 mM NaHCO3 under a nitrogen atmosphere in 150-mL anaerobic bottles. AM3 is a phosphate-based defined medium containing vitamins and the three amino acids cysteine, methionine, and glutamate (29). Pre-cultures were grown in the same medium in 10-mL volumes in 28-mL anaerobic tubes. Anaerobic tubes and bottles were flushed with O -free N gas 2 2 (Airgas, Independence, OH) for 10 min, stoppered with rubber bungs, flushed for another 10 min and sealed with aluminum crimps. For growth on glycerol, A. succinogenes was grown in AM3 supplemented with 150 mM glycerol and 150 mM NaHCO in continuous culture conditions as 3 described (30). Cultures were quenched by mixing 25-mL of culture at an OD660 of ~ 0.9-1.0 with 25-mL cold methanol and stored at -20 °C until RNA purification. Two independent 87 biological replicates were harvested for each condition. For strains expressing the synthetic sRNA constructs grown on lactose, cultures were supplemented with 50 mM lactose and 40 μg mL-1 ampicillin. All strains used in this study are listed in Table 3.1. Table 3.1 Strains and plasmids used in this study Description Source E. coli DH5α F- φ80lacZΔM15 Δ(lacZYA-argF)U169 recA1 endA1 hsdR17 Laboratory collection TOP10 (rk-, mk+) phoA supE44 λ- thi-1 gyrA96 relA1 F- mcrA Δ(mrrhsdRMS-mcrBC) φ80lacZΔM15 ΔlacΧ74 recA1 araD139 Δ(araleu) 7697 galU galK rpsL (StrR) endA1 nupG λ- Invitrogen A. succinogenes 130Z (ATCC55618) Wild-type strain ATCC Plasmids pCR2.1 TOPO AmpR, KmR, lacZα, cloning vector pLGZ920 E. coli-A. succinogenes shuttle vector; AmpR; A. succinogenes ppckA pLGZ920*-ppckA-92smRNA_lacZ1 pLGZ920 derivative containing synthetic sRNA smRNA_lacZ1 downstream of ppckA-92 promoter This study pLGZ920*-ppckA-92smRNA_lacZ2 pLGZ920 derivative containing synthetic sRNA smRNA_lacZ2 downstream of ppckA-92 promoter This study pLGZ920*-ppckA-92smRNA_lacZ3 pLGZ920 derivative containing synthetic sRNA smRNA_lacZ3 downstream of ppckA-92 promoter This study pLGZ920*-ppckA-92smRNA_lacZ4 pLGZ920 derivative containing synthetic sRNA smRNA_lacZ4 downstream of ppckA-92 promoter This study pLGZ920*-ppckA-92smRNA_ackA1 pLGZ920 derivative containing synthetic sRNA smRNA_ackA1 downstream of ppckA-92 promoter This study pLGZ920*-ppckA-92smRNA_ackA2 pLGZ920 derivative containing synthetic sRNA smRNA_ackA2 downstream of ppckA-92 promoter This study pLGZ920*-ppckA-92smRNA_pta1 pLGZ920 derivative containing synthetic sRNA smRNA_pta1 downstream of ppckA-92 promoter This study pLGZ920*-ppckA-92smRNA_pta2 pLGZ920 derivative containing synthetic sRNA smRNA_pta2 downstream of ppckA-92 promoter This study 88 Invitrogen (31) 3.2.2 RNA isolation, library preparation, and sequencing. The 50-mL bacterial suspensions in methanol were centrifuged at 4,500×g for 15 min at 4 °C. The cell pellets were used to isolate the total RNA, including sRNAs, using the miRNeasy mini kit (Qiagen, Valencia, CA). Contaminating DNA was removed using DNase I (Ambion, Austin, TX). The RNA was tested for chromosomal DNA contamination by PCR and was then cleaned up using Zymo RNA Clean & Concentrator-5 (Zymo Research, Irvine, CA). RNA quantity and quality were analyzed on a Qubit fluorometer (Invitrogen, Carlsbad, CA) and 2100 Bioanalyzer (Agilent Technologies; Santa Clara, CA), respectively. Total RNAs that passed the quality check were submitted to the Michigan State University Research Technology Support Facility for sequencing. Small RNA libraries were created using the Illumina small RNA Sample Prep Kit and were purified from 6% native polyacrylamide gels by excising fragments in the 22to-500 bp range. The sRNA libraries were quality controlled by validating them using Qubit dsDNA, Caliper LabChipGX, and Kapa qPCR assays. The libraries were pooled in equimolar amounts for multiplexed sequencing, and loaded onto one lane of an Illumina HiSeq 2500 Rapid Run flow cell (v2). Sequencing was performed in a 2×100bp paired-end format using HiSeq Rapid SBS reagents. Bases were called using Illumina Real Time Analysis (RTA) v 1.18.64. The RTA output was demultiplexed and converted to FASTQ format with Illumina Bcl2fastq v1.8.4. 3.2.3 Trimming of reads, mapping, and sRNA identification. Demultiplexed raw FASTQ files were received in paired-end format (R1 files containing left reads and R2 files containing right reads). These paired FASTQ files were trimmed to remove low-quality bases and adapters using Trimmomatic (v0.33) (32). Illumina clip was run in palindrome mode and specified to keep both reads. The sliding window was set at 4 bp, and 89 reads were discarded when the quality dropped below 20 or when read length was below 15 bp. FastQC analysis was done before and after trimming of the reads. Reads that passed filtering were aligned with the genome (Genbank accession no. NC_009655) using Rockhopper (33, 34) in the reference-based transcript assembly mode. The aligned reads were then analyzed with Rockhopper twice with different expression parameters. In run 1, the maximum number of bases between paired-end mates and the minimum expression of UTRs and ncRNAs were set at 300 and 0.5, respectively (referred to as 300_0.5 in text ahead). In run 2, the maximum number of bases between paired end mates was 300, and the minimum expression of UTRs and ncRNAs was 0.1 (referred to as 300_0.1 in text ahead). Small RNAs were also manually identified by visually inspecting the Rockhopper alignment files in IGV. Transcript start and stop coordinates were manually determined by finding the maximum expression level and extending the transcript start and stop until the expression level dropped below 10% of the maximum. 3.2.4 Small RNA analysis. All sRNA sequences were input into ARNold (Rho-independent terminator prediction program) to detect rho-independent terminators. If Rockhopper-determined sRNAs were shorter than the manually-annotated transcripts in IGV, the coordinates used were those observed in IGV. The program was run under default conditions. The free energy for the positive sequences was noted. IntaRNA (35, 36), run using default parameters, was used to identify target mRNAs for the sRNAs. RNAfold (37) was used to determine sRNA secondary structures wherever needed. 90 3.2.5 Computational prediction of sRNAs. Small RNAs were downloaded from the web-based platform SIPHT (available at http://newbio.cs.wisc.edu/sRNA/). SIPHT can search for intergenic loci that contain putative Rho-independent terminators in any of the bacterial replicons in the NCBI database. The parameters used were those mentioned by the authors (38). The sRNAs predicted for A. succinogenes were also downloaded from the Bacterial Small Regulatory RNA Database (BSRD, http://www.bac-srna.org/BSRD/taxonomyIndexNew.jsp) (39). The BSRD parameters were a maximum E-value of 1 × 10-15, TransTerm confidence value of 87%, maximum RNAMotif score of -9, FindTerm score of -10, and minimum and maximum lengths for predicted loci of 50 and 500, respectively. INFERNAL (40), RNAz (41), and Blast against Rfam (blasts the genome against previously described sRNAs)(42)—all available through the platform RNAspace (www.rnaspace.org/) (43) were also run. INFERNAL, which searches for homologs of structural RNAs in sequence databases like the Rfam 10.0 database by building covariance models (CM), was run with the default parameters. RNAz is a program that can detect structurally conserved, functional, thermodynamically stable RNA in genome-wide screens or multiple alignments. RNAz was run using the default parameters, with a probability cutoff of 0.7, slice alignments longer than 300, window size of 200, and stepsize of 50. RNAz used the annotated genomes of A. pleuropneumoniae L20 and H. influenzae Rd KW20 for alignment with the A. succinogenes genome, using BLAST with default parameters in RNAspace. In the same RNAspace run, sequence aggregation was also done using CG-seq using default parameters (i.e., score lambda parameter of 1, minimal and maximal lengths of a conserved region of 30 and 500, minimum and 91 maximum identity thresholds of 60 and 100, respectively). CRISPRs were identified in the genome using CRISPRFinder (44) with default parameters. 3.2.6 RT-PCR validation of small RNAs. First-strand cDNA was synthesized from 1 µg of total RNA with random hexamers and SuperScript II reverse transcriptase (Invitrogen, Carlsbad, CA) according to the manufacturer’s instructions. The reverse transcriptase reaction was used as the template for PCR reactions using primers designed for each sRNA candidate to be verified. As a positive control, A. succinogenes genomic DNA was used as the template for PCR. To verify the absence of genomic DNA, DNase-treated total RNA (not subjected to reverse transcriptase) was used as the template. PCR reactions were done in a 20-µL volume containing Taq polymerase, 1X Taq buffer, 50 µM of each dNTP, 1.5 mM MgCl2, and 1.25 µM of each primer. Thermocycler conditions were as follows: 95 °C for 3 min; 32 cycles of 95 °C for 30 sec, 60 °C for 30 sec and 72 °C for 1 min; 72 °C for 5 min. PCR products were visualized on a 3% agarose gel with a low molecular weight ladder (New England Biolabs, Ipswich, MA). All sRNAs were tested using cDNA synthesized from total RNA from glucose-grown cultures, except for smRNAs 47 and 126, which were tested with cDNA synthesized from RNA obtained from glucose and glycerol cultures. 3.2.7 Construction of synthetic small RNA constructs. Synthetic sRNA constructs designed to inhibit lacZ expression—smRNA_lacZ1, smRNA_lacZ2, smRNA_lacZ3, and smRNA_lacZ4, were constructed by annealing oligonucleotides CV662 and CV663, CV664 and CV665, CV666 and CV667, and CV668 and CV669 respectively (Table S3.1). Annealed oligonucleotides were used as templates and 92 amplified using primers CV677 and CV678 (smRNA_lacZ1), and primers CV677 and CV 679 (smRNA_lacZ2, smRNA_lacZ3, and smRNA_lacZ4). PCR products were cloned into pLGZ920*-pAsuc_0701-lacZ’s BamHI and SacI sites using Gibson cloning (New England Biolabs, Ipswich, MA). pLGZ920*- pAsuc_0701-lacZ is a derivative of pLGZ920 expressing lacZ, in which the pckA promoter was replaced by the promoter of Asuc_0701, and the HindIII restriction site has been removed (see chapter 2). Digesting pLGZ920*- pAsuc_0701-lacZ with BamHI and SacI removes pAsuc_0701 and lacZ. Ligation mixtures were transformed into E. coli Top10 (Invitrogen) chemically competent cells and plated on LB agar plates containing 100 µg mL-1 ampicillin. Transformant colonies were tested for the presence of the insert by colony PCR with primers CV677 and CV640, followed by plasmid sequencing with primer CV640 (GENEWIZ, Inc., South Plainfield, NJ). Positive clones were transformed into A. succinogenes strain 130Z by electroporation as described (31). Synthetic sRNA constructs targeting acetate kinase (AckA) and phosphotransacetylase (Pta) expression—smRNA_ackA1, smRNA_ackA2, smRNA_pta1, smRNA_pta2, were ordered as g-blocks from Integrated DNA Technologies (Coralville, IA) (Table S3.1). All g-blocks were digested with BamHI and XbaI and ligated into pLGZ920*-pAsuc_0701’s BamHI and XbaI sites as described above. Colonies were screened by colony PCR using primers CV677 and CV640, and plasmid inserts were verified by sequencing (GENEWIZ, Inc.). Positive clones were transformed into A. succinogenes 130Z. 3.2.8 Beta-galactosidase assays. Strain 130Z carrying constructs smRNA_lacZ1, smRNA_lacZ2, and smRNA_lacZ4 was grown in AM3 supplemented with 50 mM lactose, 150 mM NaHCO3, and 40 µg mL-1 93 ampicillin, and harvested in mid-exponential phase. The positive control was strain 130Z (pLGZ290). Cultures (200 µL) were transferred to 96-well plates and optical density at 600 nm (OD600) was recorded on a PowerWave HT microplate reader (BioTek Instruments Inc., Winooski, VT). OD600 values were used as a measure of cell biomass. One hundred µL of cultures were transferred to a new 96-well plate and lysed by adding 100 µL of Bugbuster Protein Extraction Reagent (Millipore Sigma, Billerica, MA) and incubating for 5 min. The lysate was used as the crude extract for β-galactosidase assays. β-galactosidase assays were conducted in 96-well plates in the PowerWave HT microplate reader. Reactions (200 µL) contained 102.5 µL cell extract mixed with 67.5 µL of Z-buffer (60 mM Na2HPO4 , 40 mM NaH2PO4 , 10 mM KCl, 1 mM MgSO4, 38 mM 2-mercaptoethanol, pH 7.0) and 30 µL of 2nitrophenyl β-D-galactopyranoside (4 mg mL-1 in Z-buffer). Enzyme activity was calculated using the linear slope obtained from 2-nitrophenol production recorded at 420 nm for 20 min at room temperature. Specific activity values are reported in ΔA420 min-1 ml-1 OD600-1 and are the average of three independent biological replicates. 3.2.9 Determination of fermentation balances. Growth of A. succinogenes in liquid cultures was monitored by measuring OD660 on a DU 650 spectrophotometer (Beckman, Fullerton, CA). Samples collected early (OD660 ~ 0.5) and late (OD660 ~ 1.5) in the exponential phase were used to determine the carbon balance. Glucose and organic acids were quantified in the 10-fold diluted, filtered culture supernatants by HPLC (Waters, Milford, MA) using an Aminex HPX-87H column (Bio-rad). Samples were run with 4 mM H2SO4 as the eluent at a 0.6 mL min-1 flow rate at room temperature. Glucose and ethanol were quantified on a Waters 410 differential refractometer while organic acids were quantified 94 on a Waters 2487 UV detector at 210 nm. OD660 values were used to calculate the biomass as described (45). 3.3 Results and Discussion 3.3.1 Detection of sRNAs from RNAseq data using RockHopper. The overall approach we took for identifying sRNAs is described in Figure 3.1. A. succinogenes sRNAs were sequenced in cultures grown anaerobically on glucose and microaerobically on glycerol, with two independent biological replicates for each condition. Over 95% of the read pairs remained after trimming and quality control, the details of which are mentioned in Table 3.2. Figure 3.1 Flow chart of the approach used for sRNA identification in A. succinogenes. 95 Table 3.2 Statistics on read alignments by Rockhopper Sample Glucose-1 Glucose-2 Glycerol-1 Glycerol-2 No of read Read pairs Successful Alignment to Alignment to pairs remaining after alignment by protein coding protein coding before trimming Rockhopper genes (sense) genes (antisense) trimming (% surviving) (% alignment) (%) (%) 12,495,687 13,066,506 85 16 1 (95.6%) 11,809,343 12,423,762 85 16 1 (95.0%) 10,452,076 10,842,533 86 9 2 (96.4%) 10,236,278 10,637,886 85 11 2 (96.2%) 96 Alignment to rRNAs (%) Alignment to tRNAs (%) Alignment to unannotated regions (%) 65 5 13 66 5 12 76 5 7 73 5 10 The Rockhopper-predicted sRNAs were manually inspected by visualization in IGV. Three types of Rockhopper-predicted sRNAs were discarded or modified: i) sRNAs in sense direction located in 3’UTR or 5’UTR region of annotated mRNA, whose transcript levels did not differ much from the corresponding mRNA (sense) transcript levels were deleted; ii) sRNAs that were shorter than 30 nt were deleted, unless they appeared to be longer in IGV; and iii) two or more separate RNAs that were immediately consecutive but that, upon visual inspection, seemed to belong to one bigger sRNA transcript, were listed as a unique, larger transcript. After analysis of the Rockhopper results (from 300_0.5 and 300_0.1), a total of 145 sRNAs were detected in glucose and glycerol cultures combined (Table 3. 3). Of these 145 sRNAs, five were found in glucose-grown cultures only, while 40 were found in glycerol-grown cultures only (Table 3.3). All others were found in both culture conditions. An additional 115 more sRNAs that had been missed by Rockhopper were found upon manual inspection in IGV (Table S3.1). In total, these 260 sRNAs were classified into six separate categories, as described in Figure 3.2. Figure 3.2 Classification of identified sRNAs. 97 Table 3.3 A. succinogenes sRNAs detected by RNAseq and Rockhopper analysis in anaerobically-grown glucose cultures and microaerobically-grown glycerol cultures Candidates smRNA1 smRNA2 smRNA3 smRNA4 smRNA5 smRNA6 smRNA7 smRNA8 smRNA9 smRNA10 smRNA11 smRNA12 smRNA13 smRNA14** smRNA15 smRNA16 smRNA17 smRNA18 smRNA19 smRNA20 smRNA21 smRNA22 smRNA23 smRNA24 smRNA25 smRNA26 smRNA27 Rockhopper parameters; max bases between paired end mates=300 ; minimum expression of UTRs and nc RNAS=0.5 Rfam Transcription Transcription Predicted by algorithms Size Strand Neighborhood Annotation start stop 134,044 134,063 19 + ASUC_RS00615/ASUC_RS00620 368,282 368,340 58 + ASUC_RS01695/ASUC_RS01700 369,316 369,371 55 + ASUC_RS01700/ASUC_RS01705 PyrR 431,715 431,814 99 ASUC_RS01960/ASUC_RS01965 I 517,122 517,162 40 + ASUC_RS02365/ASUC_RS02370 Z 553,139 553,170 31 ASUC_RS02590/ASUC_RS02595 609,640 609,664 24 + ASUC_RS02865/ASUC_RS02870 744,502 744,593 91 ASUC_RS03550/ASUC_RS03555 785,724 785,705 19 ASUC_RS03760/ASUC_RS03765 Z 837,968 838,075 107 ASUC_RS04020/ASUC_RS04025 MicF 981,582 981,628 46 + ASUC_RS04740/ASUC_RS04745 I 1,000,490 1,000,464 26 ASUC_RS04845/ASUC_RS04850 1,013,544 1,013,530 14 ASUC_RS04900/ASUC_RS04905 1,044,143 1,044,177 34 ASUC_RS05040/ASUC_RS05045 1,045,717 1,045,764 47 ASUC_RS05045/ASUC_RS05050 1,048,744 1,048,726 18 ASUC_RS05055/ASUC_RS05060 1,136,086 1,136,035 51 ASUC_RS05410/ASUC_RS05415 SRP_bact 1,215,624 1,215,735 111 + ASUC_RS05760/ASUC_RS05765 B, I, Z, M 1,237,630 1,237,591 39 ASUC_RS05855/ASUC_RS0560 1,276,490 1,276,511 21 + ASUC_RS06025/ASUC_RS06030 Z 1,276,603 1,276,695 92 + ASUC_RS06025/ASUC_RS06030 Z 1,367,934 1,367,964 30 + ASUC_RS06370/ASUC_RS06375 1,376,784 1,376,856 72 + ASUC_RS06445/ASUC_RS11015 Z 1,441,632 1,441,660 28 + ASUC_RS06770/ASUC_RS06775 S 1,525,386 1,525,407 21 ASUC_RS07160/ASUC_RS07165 1,545,573 1,545,663 90 + ASUC_RS07260/ASUC_RS07265 Z GcvB 1,862,491 1,862,517 26 + ASUC_RS08780/ASUC_RS08785 B, I, M 98 Table 3.3 (cont’d) smRNA28 smRNA29 smRNA30 smRNA31 2,003,880 2,003,940 60 + ASUC_RS09455/ASUC_RS09460 S 2,089,221 2,089,240 19 + ASUC_RS09840/ASUC_RS09845 Z 2,204,631 2,204,672 41 ASUC_RS10330/ASUC_RS10335 2,275,657 2,275,690 33 ASUC_RS10685/ASUC_RS10690 Rockhopper parameters; max bases between paired end mates=300 ; minimum expression of UTRs and nc RNAS=0.1 smRNA32 22 81 59 ASUC_RS10935/ASUC_RS00005 smRNA33* 16,626 16,657 31 ASUC_RS00085/ASUC_RS00090 smRNA34* 16,749 16,849 100 ASUC_RS00085/ASUC_RS00090 smRNA35* 17,297 17,329 32 ASUC_RS00090/ASUC_RS00095 smRNA36 44,059 44,089 30 + ASUC_RS00210/ASUC_RS00215 smRNA37 74,727 74,758 31 + ASUC_RS00320/ASUC_RS00325 smRNA38* 91,637 91,689 52 ASUC_RS00400/ASUC_RS00405 smRNA39 132,913 132,945 32 ASUC_RS00610/ASUC_RS00615 smRNA40 134,004 134,071 67 ASUC_RS00615/ASUC_RS00620 smRNA41 278,628 278,675 47 ASUC_RS01270/ASUC_RS01275 smRNA42 289,411 289,455 44 ASUC_RS01335/ASUC_RS01340 smRNA43 313,893 314,010 117 + ASUC_RS01445/ASUC_RS01450 smRNA44* 330,475 330,511 36 ASUC_RS01530/ASUC_RS01535 smRNA45 330,852 330,883 31 ASUC_RS01535/ASUC_RS01540 smRNA46** 353,040 353,070 30 + ASUC_RS01650/ASUC_RS01655 smRNA47** 356,657 356,723 66 ASUC_RS01655/ASUC_RS01660 smRNA48** 359,036 359,095 59 ASUC_RS01660/ASUC_RS01665 smRNA49 412,230 412,339 109 + ASUC_RS01890/ASUC_RS01895 smRNA50 501,364 501,412 48 + ASUC_RS02295/ASUC_RS02300 smRNA51 508,217 508,265 48 ASUC_RS02325/ASUC_RS02330 smRNA52 520,787 520,823 36 ASUC_RS02400/ASUC_RS02405 smRNA53 522,418 522,483 65 ASUC_RS02420/ASUC_RS02425 smRNA54 525,832 525,866 34 ASUC_RS02460/ASUC_RS02465 smRNA55 Alpha_RBS 527,930 528,046 116 + ASUC_RS02475/ASUC_RS02480 S, B, I, Z, M smRNA56 596,463 596,584 121 ASUC_RS02810/ASUC_RS02825 smRNA57 Parecho_CRE 599,496 599,569 73 + ASUC_RS02835/ASUC_RS02840 I 99 Table 3.3 (cont’d) smRNA58* smRNA59* smRNA60 smRNA61* smRNA62* smRNA63 smRNA64 smRNA65** smRNA66* smRNA67 smRNA68 smRNA69* smRNA70* smRNA71* smRNA72* smRNA73 smRNA74* smRNA75 smRNA76 smRNA77 smRNA78 smRNA79 smRNA80 smRNA81 smRNA82* smRNA83* smRNA84 smRNA85 smRNA86* smRNA87 Glycine riboswitch Thr_leader Parecho_CRE 604,037 604,515 620,650 632,299 632,547 643,335 643,644 656,185 657,344 694,517 705,794 743,678 744,115 744,307 744,450 769,202 785,488 787,773 604,135 604,583 620,681 632,398 632,583 643,371 643,675 656,204 657,404 694,526 705,810 743,708 744,089 744,380 744,438 769,245 785,512 787,816 98 68 31 99 36 36 31 19 60 9 16 30 26 73 12 43 24 43 + + + + + + + + + + + ASUC_RS02845/ASUC_RS02850 ASUC_RS02845/ASUC_RS02850 ASUC_RS02915/ASUC_RS02920 ASUC_RS02980/ASUC_RS02985 ASUC_RS02980/ASUC_RS02985 ASUC_RS03035/ASUC_RS03040 ASUC_RS03035/ASUC_RS03040 ASUC_RS03080/ASUC_RS03085 ASUC_RS03085/ASUC_RS03090 ASUC_RS03285/ASUC_RS03290 ASUC_RS03340/ASUC_RS03345 ASUC_RS03545/ASUC_RS03550 ASUC_RS03545/ASUC_RS03550 ASUC_RS03550/ASUC_RS03555 ASUC_RS03550/ASUC_RS03555 ASUC_RS03670/ASUC_RS03680 ASUC_RS03755/ASUC_RS03760 ASUC_RS03770/ASUC_RS03775 806,342 806,566 224 + ASUC_RS03870/ASUC_RS03875 B, I 858,732 869,489 871,031 883,325 906,434 924,096 926,654 936,865 938,190 1,012,701 1,014,494 858,775 869,546 871,049 883,345 906,465 924,135 926,690 936,906 938,323 1,012,743 1,014,571 43 57 18 20 31 39 36 41 133 42 77 + + + + + + + + + ASUC_RS04140/ASUC_RS04145 ASUC_RS04185/ASUC_RS04190 ASUC_RS04190/ASUC_RS04195 ASUC_RS04250/ASUC_RS04255 ASUC_RS04380/ASUC_RS04385 ASUC_RS04460/ASUC_RS04465 ASUC_RS04465/ASUC_RS04470 ASUC_RS04530/ASUC_RS04535 ASUC_RS04535/ASUC_RS04550 ASUC_RS04895/ASUC_RS04900 ASUC_RS04905/ASUC_RS04910 M I 100 Table 3.3 (cont’d) smRNA88 smRNA89* smRNA90* smRNA91 smRNA92 smRNA93 smRNA94 smRNA95* smRNA96 smRNA97* smRNA98 smRNA99* smRNA100* smRNA101* smRNA102* smRNA103 smRNA104* smRNA105 smRNA106* smRNA107* smRNA108 smRNA109 smRNA110 smRNA111* smRNA112 smRNA113 smRNA114 smRNA115 smRNA116* smRNA117 smRNA118 1,057,772 1,067,837 1,069,547 1,073,915 1,083,430 1,117,501 1,119,016 1,146,049 1,165,100 1,240,529 1,253,225 1,250,832 1,252,103 1,252,811 1,264,237 1,265,927 1,266,017 1,266,795 1,267,226 1,267,779 1,296,921 1,298,854 1,300,630 1,368,446 1,370,455 1,376,675 1,377,939 1,400,489 1,415,979 1,424,398 1,455,834 1,057,841 1,067,870 1,069,772 1,073,974 1,083,410 1,117,539 1,119,057 1,146,091 1,165,133 1,240,560 1,253,175 1,250,866 1,252,198 1,252,842 1,264,348 1,265,963 1,266,073 1,266,826 1,267,256 1,267,810 1,296,999 1,298,937 1,300,662 1,368,552 1,370,499 1,376,702 1,377,967 1,400,682 1,416,016 1,424,495 1,455,884 69 33 225 59 20 38 41 42 33 31 50 34 95 31 111 36 56 31 30 31 78 83 32 106 44 27 28 193 37 97 50 + + + + + + + + + + + + + + + + + ASUC_RS05080/ASUC_RS05085 ASUC_RS05120/ASUC_RS05125 ASUC_RS05125/ASUC_RS05130 ASUC_RS05155/ASUC_RS05160 ASUC_RS05215/ASUC_RS05220 ASUC_RS05330/ASUC_RS05335 ASUC_RS05330/ASUC_RS05340 ASUC_RS05450/ASUC_RS05455 ASUC_RS05545/ASUC_RS05550 ASUC_RS05865/ASUC_RS05870 ASUC_RS05925/ASUC_RS05930 ASUC_RS05915/ASUC_RS05920 ASUC_RS05920/ASUC_RS05925 ASUC_RS05925/ASUC_RS05930 ASUC_RS05980/ASUC_RS05985 ASUC_RS05990/ASUC_RS05995 ASUC_RS05990/ASUC_RS05995 ASUC_RS05990/ASUC_RS05995 ASUC_RS05990/ASUC_RS05995 ASUC_RS05995/ASUC_RS06000 ASUC_RS06085/ASUC_RS06090 ASUC_RS06085/ASUC_RS06090 ASUC_RS06085/ASUC_RS06090 ASUC_RS06370/ASUC_RS06375 ASUC_RS06390/ASUC_RS06395 ASUC_RS06445/ASUC_RS11015 ASUC_RS11015/ASUC_RS06455 ASUC_RS06585/ASUC_RS06590 ASUC_RS06650/ASUC_RS06655 ASUC_RS06685/ASUC_RS06690 ASUC_RS06840/ASUC_RS06845 101 S Table 3.3 (cont’d) smRNA119* 1,493,409 1,493,453 44 + ASUC_RS07025/ASUC_RS07030 smRNA120 1,499,794 1,499,829 35 ASUC_RS07050/ASUC_RS07055 smRNA121 1,528,991 1,528,963 28 ASUC_RS07175/ASUC_RS07180 smRNA122 1,543,327 1,543,374 47 + ASUC_RS07245/ASUC_RS07250 smRNA123 1,586,321 1,586,358 37 ASUC_RS07425/ASUC_RS07430 smRNA124 1,623,706 1,623,739 33 + ASUC_RS07620/ASUC_RS07625 smRNA125 1,671,194 1,671,207 13 + ASUC_RS07840/ASUC_RS07845 smRNA126* 1,677,571 1,677,675 104 ASUC_RS07870/ASUC_RS07875 smRNA127 1,700,597 1,700,770 173 ASUC_RS07990/ASUC_RS07995 smRNA128 1,737,609 1,737,650 41 ASUC_RS08165/ASUC_RS08170 smRNA129* 1,901,859 1,901,893 34 ASUC_RS08940/ASUC_RS08945 smRNA130* 1,902,651 1,902,682 31 ASUC_RS08945/ASUC_RS08950 smRNA131* 1,927,500 1,927,538 38 ASUC_RS09070/ASUC_RS09075 smRNA132 1,974,287 1,974,272 15 ASUC_RS09300/ASUC_RS09305 smRNA133 1,977,589 1,977,719 130 ASUC_RS09310/ASUC_RS09315 smRNA134* 1,978,600 1,978,637 37 + ASUC_RS09320/ASUC_RS09325 smRNA135 1,981,750 1,981,830 80 ASUC_RS09340/ASUC_RS09345 smRNA136 1,989,202 1,989,252 50 ASUC_RS09380/ASUC_RS09385 smRNA137 2,099,513 2,099,588 75 + ASUC_RS09885/ASUC_RS09890 smRNA138 2,202,223 2,202,295 72 + ASUC_RS10320/ASUC_RS10325 smRNA139 2,220,187 2,220,167 20 ASUC_RS10410/ASUC_RS10415 smRNA140 2,266,375 2,266,447 72 ASUC_RS10650/ASUC_RS10655 smRNA141 2,291,971 2,291,996 25 + ASUC_RS10785/ASUC_RS10790 smRNA142* 2,300,851 2,300,963 112 + ASUC_RS10825/ASUC_RS10830 smRNA143* 2,307,624 2,307,693 69 ASUC_RS10870/ASUC_RS10875 smRNA144 2,314,470 2,314,449 21 ASUC_RS10900/ASUC_RS10905 smRNA145* 2,315,613 2,315,645 32 ASUC_RS10910/ASUC_RS10915 **smRNA expressed in glucose cultures only * smRNA expressed in glycerol cultures only. A minimum cutoff of 50 for transcript level was used to determine whether or not a sRNA was expressed. B, BSRD; I, INFERNAL; M, Rfam; S, SIPHT; Z, RNAz 102 Z 5’UTRs were defined as transcripts whose 3’ end preceded a coding sequence by fewer than 50 nt. In some cases riboswitch transcripts did not end until well within the coding sequence. For ease of classification, these riboswitches were still grouped in the 5’UTR category. 3’UTRs were defined as transcripts whose 5’ end started within the coding region or immediately after the stop codon. Many transcripts started well within the coding sequence (sense direction) and extended beyond the coding sequence, usually accompanied by lower transcript levels of the gene itself (Figure 3.3). Antisense sRNAs were divided into three categories: as5’UTRs, cis-asRNAs, and as3’UTRs. Transcripts antisense to the 5’UTR of the coding sequence that started within 50 nt of the start codon were called as5’UTRs. Cis-asRNAs were defined as antisense and started within the coding sequence. Figure 3.3 Example of 3’UTR sRNA overlapping with the mRNA transcript and downstream intergenic region with both mRNA and sRNA encoded on the reverse strand. -, reverse strand; +, forward strand; red: reads on the reverse strand; blue: reads on the forward strand; white arrowheads in gene indicate direction of the gene. Figure modified from screenshot in IGV. Transcripts antisense to the coding sequence that started within 50 nt of the stop codon were called as3’UTRs. All others were called trans-encoded sRNAs. Using this classification, 103 the 260 sRNAs were catalogued as 28 5’UTRs, 71 3’UTRs, 2 as5’UTRs, 81 asRNAs, 10 as3’UTRs, and 68 trans-encoded sRNAs. 3.3.2 Comparison of RNAseq-identified to computationally predicted sRNAs. SIPHT predicted 45 sRNAs, of which 40 were unknown sRNAs and five were known sRNAs or riboswitches annotated in Rfam. Eleven of the SIPHT-predicted sRNAs were also identified by sRNAseq, but some of them were combined since they looked like a single transcript in IGV, giving a total of six sRNAs shared between the two—smRNA24, smRNA28, smRNA55, smRNA88, smRNA175, and smRNA251 (Table S3.2). SIPHT predicted two sRNAs that were already annotated in A. succinogenes as rRNAs or tRNAs—asuc_R0067 and asuc_RS0071 (NCBI old locus tags). These were omitted in our analysis. For smRNA24, SIPHT predicted nine different transcripts, five of which started from the same coordinate but had different stop coordinates. The other four all had different start coordinates but shared the same stop coordinate. Because smRNA24 corresponded to a continuous transcript in IGV, it was considered to be a single transcript. SIPHT also predicted smRNA175 as two separate transcripts, but since IGV showed it as a continuous transcript, it was deemed to be one sRNA. Two of the sRNAs predicted by SIPHT and detected in RNAseq are already annotated in Rfam as Alpha_RBS (smRNA55) and CRISPR-DR34 (smRNA175) while the other four are unknown. BSRD (Bacterial Small Regulatory RNA Database) lists eighteen sRNA candidates for A. succinogenes 130Z, all of which are orthologs of known sRNAs in other bacterial species. Twelve of these were detected in our RNAseq study, of which three (not listed in Table S3.3) are already annotated in A. succinogenes: asuc_R0067 (NCBI old locus tag) (annotated as tmRNA in Rfam), ASUC_RS11040 (annotated as Rnase P classA in Rfam), and ASUC_RS09385 104 (annotated as S15 in Rfam). These three sRNAs were not considered for further analysis. The remaining nine are listed in Table S3.3. BSRD lists two separate sRNAs spanning coordinates 806377‒806453 and 806464‒806566. Rockhopper predicted a single sRNA instead (smRNA76), with coordinates 806342‒806566. smRNA76 was considered to be one single sRNA after visual confirmation in IGV (Figure 3.4). Figure 3.4 smRNA76 is seen as a single transcript in IGV. +, forward strand; -, reverse strand; blue: reads detected in RNAseq; green: Rockhopper-predicted sRNA in both glucose and glycerol libraries; red bars, two transcripts predicted by BSRD. Figure modified from screenshot in IGV. INFERNAL, RNAz, and Blastn against Rfam predicted 167, 108, and 115 putative non coding RNAs (including tRNAs and rRNAs) respectively. Fifteen were common between INFERNAL and RNA-seq, all of which are annotated in Rfam (Table S3.4). INFERNAL predicted two separate transcripts for smRNA76, which we described above as a single transcript (Figure 3.4 and Table S3.4). Twenty sRNAs were common between RNAz and RNA-seq, three of which are annotated in Rfam (Table S3.5). RNAseq and Blastn against Rfam had ten sRNAs in common, with annotations listed in Table S3.6. CRISPRFinder predicted two CRISPR elements in the A. 105 succinogenes genome, which were both detected in RNAseq. One of the CRISPR elements is annotated in Rfam as CRISPR –DR34 (smRNA175) (predicted as 7 separate trascripts in Rfam spanning coordinates 436,472 and 436,903), while the other is unannotated (smRNA24). smRNA24 had previously been described in our A. succinogenes genome paper (46) using CRISPRs web service. Overall, thirteen of small RNAs detected in RNAseq were also detected by at least one of the computational programs. Twenty-nine of the sRNAs detected by RNAseq were predicted by at least one of the computational programs, three were predicted by two computational programs, four were predicted by three of the programs, and two were predicted by four programs, while only one was predicted by all five computational programs. 3.3.3 sRNAs with Rho-independent terminators. Many sRNAs depend on Hfq, a chaperone protein, to bind to their target mRNAs and for stability (4). Hfq typically binds to the U-rich region immediately preceding the poly(U) tail found in RNAs with Rho-independent terminators. Thus sRNAs with Rho-independent terminators are possibly Hfq-dependent. Twenty four Rho-independent terminators were identified among the 260 sRNAs detected in RNAseq (Rockhopper and manually identified) using ARNold (Table 3.4). 3.3.4 RT-PCR validation of small RNAs. We selected fourteen sRNAs for validation with RT-PCR—smRNA4, 8, 11, 18, 21, 27, 28, 47, 49, 55, 76, 126, 135, and 187. smRNA4 was selected because it was predicted by a single computational algorithm, INFERNAL. smRNA8 and smRNA27 were not predicted by any of the computational algorithms. smRNA11 was selected because it is annotated as MicF in Rfam and was also predicted by INFERNAL. smRNA18 was selected because it was antisense to ffs 106 Table 3.4 Rho-independent smRNAs predicted by ARNold sRNA candidate Transcription Transcription starta stopa Size Strand Neighborhood Free energyb (kcal/mol ) smRNA6 553,132 553,188 56 - ASUC_RS02590/ASUC_RS02595 -5.06 smRNA8 744,498 744,593 95 - ASUC_RS03550/ASUC_RS03555 -4.80 smRNA11 981,582 981,635 53 + ASUC_RS04740/ASUC_RS04745 -12.7 smRNA13 1,013,483 1,013,544 61 - ASUC_RS04900/ASUC_RS04905 -11.1 smRNA15 1,045,695 1,045,765 70 - ASUC_RS05045/ASUC_RS05050 -8.45 smRNA16 1,048,685 1,048,748 63 - ASUC_RS05055/ASUC_RS05060 -12.26 smRNA17 1,135,901 1,136,086 185 - ASUC_RS05410/ASUC_RS05415 -11.2 smRNA25 1,525,370 1,525,415 45 - ASUC_RS07160/ASUC_RS07165 -12.3 smRNA27 1,862,486 1,862,541 55 + ASUC_RS08780/ASUC_RS08785 -12.86 smRNA28 2,003,880 2,003,947 67 + ASUC_RS09455/ASUC_RS09460 -8.8 smRNA31 2,275,649 2,275,708 59 - ASUC_RS10685/ASUC_RS10690 -10.67 smRNA77 858,732 858,865 133 + ASUC_RS04140/ASUC_RS04145 -13.1 smRNA79 871,031 871,079 48 + ASUC_RS04190/ASUC_RS04195 -9.3 smRNA88 1,057,768 1,057,841 73 - ASUC_RS05080/ASUC_RS05085 -11.1 smRNA113 1,376,675 1,376,719 44 + ASUC_RS06445/ASUC_RS11015 -3.4 smRNA132 1,974,240 1,974,280 40 - ASUC_RS09300/ASUC_RS09305 -9.2 smRNA153 102,498 102,589 91 + ASUC_RS003455/ASUC_RS00460 -11.64 smRNA175c 436,421 436,943 522 - ASUC_RS01975/ASUC_RS01980 -10.4 smRNA185 691,158 691,339 181 + ASUC_RS03270/ASUC_RS03275 -8.99 smRNA207 1,169,658 1,169,834 176 + ASUC_RS05560/ASUC_RS05565 -11.1 smRNA209 1,185,693 1,185,896 203 - ASUC_RS05625/ASUC_RS05630 -11.2 smRNA227 1,520,160 1,520,316 156 - ASUC_RS07125/ASUC_RS07130 -14.2 smRNA238 1,724,668 1,724,828 160 - ASUC_RS08120/ASUC_RS08125 -10.3 smRNA240 1,825,558 1,825,646 88 ASUC_RS08605/ASUC_RS08610 -7.5 sRNA sizes predicted by Rockhopper were often shorter than the transcripts seen in IGV and did not include the rho-independent terminators. Rockhopper coordinates were extended at the 3’end until they included the rhoindependent terminator, but only until the transcript ends seen in IGV. These extended corodinates were used as input in ARNold and are listed in this table. a b Free energy of the predicted terminator stem-loop structure using RNAfold (47), can be used as a confidence value for predicted terminators. c Rho-independent terminator is also predicted about 265 nts into the transcript of smRNA175, which is annotated as CRISPR-DR34 with a free energy of stem loop region of -6.10 kcal/mol. (encoding the signal recognition particle 4.5S RNA), was predicted by most algorithms, and is annotated as SRP_bact in Rfam (Table 3.2). smRNA21 and smRNA137 were only predicted by RNAz. smRNA27 is annotated as GcvB in Rfam and was selected because it was identified as a 107 regulatory ncRNA in E.coli and other enteric bacteria. It regulates several protein-encoding genes involved in metabolic pathways (48). smRNA47 was detected only in the glucose-grown cultures in our RNAseq analysis (any sRNA expression level below 50 was considered as not expressed). smRNA49 was selected as it was classified as a 3’UTR sRNA in our analysis. smRNA55 was classified as a 5’UTR and annotated in Rfam as Alpha_RBS. smRNA76 was selected because BSRD predicts it as 2 separate transcripts, but it looks like a single transcript on IGV (Figure 3.3). smRNA126 was classified as an as3’UTR and was detected only in the glycerol cultures. smRNA187 was selected as it is annotated as 6S RNA in Rfam. All sRNAs tested by RT-PCR showed a PCR product of the expected size (Figure 3.5). smRNA47 was tested with cDNA synthesized from RNA extracted from both growth conditions. Even though smRNA47 was not detected in our RNAseq data of glycerol cultures, it produced a PCR band in both conditions. This may be due to a small amount of transcript being expressed in AM3-glycerol that may have been missed by RNAseq, or the expression level was below our cutoff of 50. Note that RT-PCR is not quantitative and that even trace amounts of DNA can be amplified to produce intense bands. It is not due to contaminating genomic DNA in the sample, as the negative control with the RNA sample does not produce a PCR band (lane 11). (Note that the band seen at ~50 bp in lane 11 [DNase-treated total RNA used as template] is likely due to primer dimers, as the same band is not seen with another primer set tested with the same template [lane 17]). A similar result (a PCR product of the expected size for both glucose and glycerol samples, lanes 15-16), with a similar interpretation, was obtained for smRNA126, for which we expected a PCR product only with the cDNA originating from the glycerol sample. For smRNA76, the primers were designed to amplify a product only if the two sRNAs predicted 108 by BSRD corresponded to a single transcript. The 240 bp PCR product seen for smRNA76 (lane 14) confirms our RNAseq results. As described above, we chose the 14 small RNAs to cover a wide variety of sRNAs belonging to different categories in our analysis. We were able to confirm all these sRNAs by RT-PCR. Based on these results, most other sRNAs detected by RNAseq are likely to be valid sRNAs. Figure 3.5 RT-PCR validation of small RNAs. Lane 1, low molecular weight ladder; Lane 2, smRNA4; Lane 3, smRNA8; Lane 4, smRNA11; Lane 5, smRNA18; Lane 6, smRNA21; Lane 7, smRNA27, Lane 8, smRNA28; Lane 9, smRNA47 (cDNA from Glucose-1); Lane 10, smRNA47 (cDNA from Glycerol-1), Lane 11, smRNA47 (DNAse treated RNA control from Glycerol-1); Lane 12, smRNA49; Lane 13, smRNA55; Lane 14, smRNA76; Lane 15, smRNA126 (cDNA from Glucose-1); Lane 16, smRNA126 (cDNA from Glycerol-1); Lane 17, smRNA126 (DNase treated RNA control from Glucose-1); Lane 18, smRNA135; Lane 19, smRNA187. 109 3.3.5 Scaffold selection for synthetic sRNA design based on predicted Hfq-dependent small RNAs. Among the twenty four Hfq-dependent small RNAs predicted by ARNold, we short-listed a few candidates with other characteristic features of Hfq-dependent small RNAs known to be important for binding to Hfq: (i) Rho-independent terminator with a long poly(U) tail, (ii) internal hairpin, and (iii) U-rich sequence preceding the hairpin (49, 50). Seven candidates (Table 3.5) met the three criteria. Out of these, smRNA8 and smRNA28 were selected as scaffolds for synthetic small RNA design. In E. coli, sRNA MicA (51) is located near the 5’ end of the luxS gene in the antisense orientation. In A. succinogenes, smRNA8 is similarly positioned upstream of luxS (ASUC_RS03555) in the opposite orientation (Figure 3.6). Although these two sRNAs do not share any sequence identity, they may be functionally similar. When we searched for target genes for smRNA8, OmpA was one of the top predicted targets (data not shown). In E. coli, MicA has been confirmed to be an antisense OmpA regulator and Hfq-dependent. We chose the smRNA8 scaffold for testing our synthetic small RNA for these reasons. We chose smRNA28 as a second scaffold, because it is conserved in at least three other Pasteurellaceae species—Mannheimia succiniciproducens, Haemophilus somnus 2336, and Haemophilus somnus 129PT. 110 Table 3.5 smRNA sequences that fit the criteria for Hfq-dependent sRNAs sRNA candidates smRNA8 (-) smRNA11 (+) smRNA13 (-) smRNA16 (-) smRNA27 (+) smRNA28 (+) smRNA77 (+) Sequences GCAGUUGUGAUUAAUAAUAAAAAAAUUGGUUCCUUAGGUUAUAUUCACCGCUCAAUUCCGCA AGGAAAAGAGCGGUUUUUUUU UUAUUCAUAAAACCCCUUUUUACGCCGGAUUCCCUUAGUCCGGCUUUUUU UGGCUAAACCGAUUUCACCCAGUAAUUCGCCGGCAUUAGUACCGACAAUCGCACCACCUAAA AUACGGUGCGAUUCUUUGUC UCCUAAAUAAAAUAAUCAUAAAUAAAUAGCUAAACUACUUUCCUACCGUCCUUUUGGACGGU UUUUUUUC UUUCUAGUUUGUCCGCUCUGCUUUCUUUUUCUACAAUACGCGCAUACUUAAUGACUGGUAAU UCCUUAAUUGAUUAAGAGUUGAAUCUUUUAGUUAAGUAUUAUGUUGUGUUUGCAUAUUGUUU GGGUAACCAAACAAAAGUAAUUAAUCCUUCUAUUUAAUUACUUAUUAACUUCCUGUAUAUUU ACUACCUAAUUUUAGGUUAUUGGCACCGCGCUUAAACUCCAAAAAAGUGCGGUGUUUUUU AUUCAAACAAUAAUAGAUAAUCACUCCAACUUUCGGCGUUUCUCUCCCCACAAGGAAACGCC UUUUUCU AUUCGUUUUUAACGGAAAAAACACCAUGAAAUCCGACCGCACUUUUACCAUGAUGACGAUUA CCACCAUUAUGACCUUUAUAAUGGCGGGGUAGUGCGAACGAAAGAUAUUCAAUUAAGCCCGC AACCUGAAAAGUGCGGGCUUUUUU + and – signs indicate the coding strand. Blue: sequences that may be complementary to target mRNA(s); red: inverted repeats in the hairpin; underline: AU-rich regions before the inverted repeats; and black and italics: poly(U) tails. Figure 3.6 A. succinogenes smRNA8’s genomic locus (Glucose-1 sample shown as a representative sample). -, reverse strand; red, transcript level from RNAseq; green, sRNA predicted by Rockhopper (in both glucose and glycerol libraries). 3.3.6 Testing of synthetic sRNA designs with lacZ as target gene. Yoo et al. (52) designed a synthetic sRNA in E. coli using the scaffold from MicC, which is a well-characterized E. coli sRNA. Yoo et al. simply changed the complementary target gene binding regions. Addition of a terminator sequence was optional since the scaffold sequence has 111 a hairpin at the 3’ end. In that study, a target binding sequence of 20-30 nt with binding energy between -30 and -40 kcal mol-1 worked best (52). Using these guidelines, we constructed four synthetic sRNA constructs targeting lacZ using smRNA8 and smRNA28 as scaffolds (Figure 3.7). smRNA_lacZ1 contains the smRNA8 scaffold sequence. smRNA_lacZ2 contains the smRNA28 scaffold sequence with a region complementary to the initial target mRNA kept intact (in blue) (Figure 3.7). smRNA_lacZ3 contains the smRNA28 scaffold sequence without any region complementary to the initial target mRNA. smRNA_lacZ1, smRNA_lacZ2, and smRNA_lacZ3 contain the same lacZ target sequence. smRNA_lacZ4 has the same scaffold as smRNA_lacZ3, but has a different lacZ target binding region. All constructs were cloned under control of a promoter of moderate strength (constructed in chapter 2). We were unable to obtain A. succinogenes transformants with pLGZ920-ppckA-92smRNA_lacZ3. Strains 130Z (pLGZ920-ppckA-92-smRNA_lacZ1), 130Z (pLGZ920-ppckA-92smRNA_lacZ2), and 130Z (pLGZ920-ppckA-92-smRNA_lacZ4) were grown on lactose to determine whether or not lacZ was expressed. Although all three strains were able to grow on AM3-lactose, 130Z (pLGZ920-ppckA-92-smRNA_lacZ1) and 130Z (pLGZ920-ppckA-92smRNA_lacZ2) grew slower than 130Z (pLGZ920-ppckA-92-smRNA_lacZ4) and 130Z (pLGZ920), the positive control. Crude extracts of strain 130Z (pLGZ920-ppckA-92smRNA_lacZ1) had a 32% decrease in β-galactosidase activity compared to the positive control (p = 0.007) (Figure 3.8). Crude extracts of strain 130Z (pLGZ920-ppckA-92-smRNA_lacZ2) showed a slight decrease in β-galactosidase activity that was not significant (p = 0.12), and βgalactosidase activity of strain 130Z pLGZ920-ppckA-92-smRNA_lacZ4) was not affected. 112 CGAAAAUUUGAUCUAGUUAACAUUUUUUAGGUAUAAAUAGUUUUAAAAUAGAUCUAGUUUGGAUUCAAAGUAGGUUGGCAGAAUCA UUCAAUUCCGCAAGGAAAAGAGCGGUUUUUUUU (A) CGAAAAUUUGAUCUAGUUAACAUUUUUUAGGUAUAAAUAGUUUUAAAAUAGAUCUAGUUUGGAUUCAAAGUAGGUUGGCAGAAUCA UAAUCACUCCAACUUUCGGCGUUUCUCUCCCCACAAGGAAACGCCUUUUUCU (B) CGAAAAUUUGAUCUAGUUAACAUUUUUUAGGUAUAAAUAGUUUUAAAAUAGAUCUAGUUUGGAUUCAAAGUAGGUUGGCAGAAUCA UAACUUUCGGCGUUUCUCUCCCCACAAGGAAACGCCUUUUUCU (C) CGAAAAUUUGAUCUAGUUAACAUUUUUUAGGUAUAAAUAGUUUUAAAAUAGAUCUAGUUUGGAUAGAAUCAUGCUGAACUCCUUAA ACUUUCGGCGUUUCUCUCCCCACAAGGAAACGCCUUUUUCU (D) lacZ target binding regions 5’ AUGAUUCUGCCAACCUACUUUGAA 3’ (E) 5’ UAAGGAGUUCAGCAUGAUUCU 3’ (F) Figure 3.7 Synthetic sRNAs targeting lacZ expression. (A) smRNA_lacZ1; (B) smRNA_lacZ2; (C) smRNA_lacZ3; (D) smRNA_lacZ4; (E) Target binding region in synthetic small RNAs in constructs A, B, and C; (F) Target binding region in synthetic sRNA construct D. Underlined:ppckA-92 promoter; green highlight: sequence complementary to lacZ mRNA; italics: scaffold sequence from smRNA8 (construct A) and smRNA28 (constructs B, C, and D); blue: sequences that may be complementary to initial target mRNA(s); red: inverted repeats preceded by AU-rich region and followed by poly(U) tail; bold: start codons in panels E and F. Binding energy for complementary regions is -36 kcal mol-1 for constructs A, B, and C, and -32.4 kcal mol-1 for construct D. ΔA420 min-1 ml-1 OD600-1 800 700 600 ** 500 400 300 200 100 0 1 2 3 4 Figure 3.8 β-galactosidase activity of strain 130Z expressing synthetic sRNAs targeting lacZ expression. 1, 130Z (pLGZ920), positive control; 2, 130Z (pLGZ920-ppckA-92-smRNA_lacZ1); 3, pLGZ920-ppckA-92-smRNA_lacZ2; and 4, pLGZ920-ppckA-92-smRNA_lacZ4. Results are the average ± standard deviation based on three independent biological replicates. 113 3.3.7 Synthetic sRNAs for inhibiting acetate production. Since a 32% decrease in β-galactosidase activity was observed using a synthetic sRNA based on the smRNA8 scaffold sequence, we used the same scaffold to produce synthetic sRNAs and decrease expression of the ackA-pta genes. The ackA and pta genes encode acetate kinase and phosphate acetyltransferase, respectively, which are responsible for acetate production from acetyl-CoA. We also used the smRNA28 scaffold of the smRNA_lacZ2 construct, which also caused a slight decrease in β-galactosidase activity. In both scaffolds, we replaced the lacZ target binding region specific for lacZ with ackA and pta target binding regions (Figure 3.9) to generate four constructs. Fermentation balances of strains 130Z (pLGZ920-ppckA-92-smRNA_ackA1), 130Z (pLGZ920-ppckA-92-smRNA_ackA2), 130Z (pLGZ920- ppckA-92-smRNA_pta1) and 130Z (pLGZ920-ppckA-92-smRNA_pta2) grown anaerobically on glucose are shown in Table 3.6. Only strain 130Z (pLGZ920-ppckA-92-smRNA_ackA2) showed a significant decrease, 14%, in acetate production compared to 130Z (pLGZ920). Surprisingly, the two sRNA constructs based on the smRNA28 scaffold caused significant increases in acetate and succinate production at the expense of ethanol and biomass production. 114 CGAAAAUUUGAUCUAGUUAACAUUUUUUAGGUAUAAAUAGUUUUAAAAUAGAUCUAGUUUGGACAGUUAAGAAUUAAAACUAAU UUGGACAUUCAAUUCCGCAAGGAAAAGAGCGGUUUUUUUU (A) CGAAAAUUUGAUCUAGUUAACAUUUUUUAGGUAUAAAUAGUUUUAAAAUAGAUCUAGUUUGGACAGUUAAGAAUUAAAACUAAU UUGGACAUAAUCACUCCAACUUUCGGCGUUUCUCUCCCCACAAGGAAACGCCUUUUUCU (B) CGAAAAUUUGAUCUAGUUAACAUUUUUUAGGUAUAAAUAGUUUUAAAAUAGAUCUAGUUUGGAGGGAUAAGAAUAAAUGUACGAGA CAUUCAAUUCCGCAAGGAAAAGAGCGGUUUUUUUU (C) CGAAAAUUUGAUCUAGUUAACAUUUUUUAGGUAUAAAUAGUUUUAAAAUAGAUCUAGUUUGGAGGGAUAAGAAUAAAUGUACGAGA CAUAAUCACUCCAACUUUCGGCGUUUCUCUCCCCACAAGGAAACGCCUUUUUCU (D) ackA target binding region 5’ AUGUCCAAAUUAGUUUUAAUUCUUAACUG 3’ (E) pta target binding region 5’ AUGUCUCGUACAUUUAUUCUUAUCCC 3’ (F) Figure 3.9 Synthetic sRNA designs targeting expression of ackA and pta. (A) smRNA_ackA1; (B) smRNA_ackA2; (C) smRNA_pta1; (D) smRNApta2; (E) ackA target binding regions in synthetic sRNAs A and B; (F) pta target binding regions in synthetic sRNAs C and D. Underlined: ppckA-92 promoter; green highlight: sequences complementary to ackA mRNA (A and B) and to pta mRNA (C and D) (binding energy is -35.2 kcal mol-1 [A and B] and -36.7 kcal mol-1 [C and D]); italics: scaffold sequence from smRNA_lacZ1 (A and C) and smRNA_lacZ2 (B and D); red: inverted repeats preceded by AU-rich region and followed by poly(U) tail; bold: start codons. 115 Table 3.6 Fermentation balances of strain 130Z expressing synthetic sRNA constructs. Products (mmol/100 mmol glucose consumed) Strain Succinate 130Z (pLGZ920) 47.9 ± 1.4 130Z (pLGZ920-ppckA-92-smRNA_ackA1) 58.9 ± 4.2* 130Z (pLGZ920-ppckA-92-smRNA_ackA2) 46.5 ± 1.4 130Z (pLGZ920-ppckA-92-smRNA_pta1) 60.8 ± 4.2* 130Z (pLGZ920-ppckA-92-smRNA_pta2) 56.6 ± 5.4 Formate 100 ± 2 106 ± 10 81.1 ± 4.2 103 ± 10 97.2 ± 8.8 Acetate 65.5 ± 1.6 76.8 ± 6.5* 56.1 ± 2.2** 73.6 ± 3.8* 68.2 ± 5.4 Ethanol 36 ± 3.4 25.3 ± 3.6 36.5 ± 16.9 25.5 ± 12.5 22.9 ± 7.6 CO2 2.06 ± 1.87 0.0±0.0 0.0±0.0 0.0±0.0 0.0 ±0.0 Biomassa 193 ± 10 165 ± 8 186 ± 3 146 ± 3 149 ± 8 Carbon recoveryb 104 ± 6 104 ± 1 105 ± 3 102 ± 2 100 ± 9 Doubling time 2.05 ± 0.03 2.45 ± 0.15 2.17 ± 0.05 2.51 ± 0.01 2.40 ± 0.09 Results are an average of three biological replicates ± standard deviations. a Biomass was determined using assumed values of 567 mg dry cell weight/mL per OD660 (45) and a cell composition of CH2O0.5N0.2 (24.967 g/mol) (53) b Carbon balance is the carbon in products/carbon in glucose consumed. It is assumed that one CO2 is fixed for each molecule of succinate produced. CO2 is calculated using the following formula: CO2 (in mM)= Ethanol (in mM)+ Acetate (in mM) - Formate (in mM) * Significantly different from 130Z (pLGZ920) (p < 0.05, two-tailed student’s t-test) ** Significantly different from 130Z (pLGZ920) (p < 0.01, two-tailed student’s t-test) 116 3.4 Conclusion In this study, we have successfully identified sRNAs in A. succinogenes and verified select ones using RT-PCR. In all, 260 sRNAs were identified using Rockhopper and manual annotation using IGV, in A. succinogenes grown anaerobically on glucose and/or microaerobically on glycerol. Combined with RNAseq, we used five computational programs to identify additional small RNAs in A. succinogenes. Thirty-nine of the sRNAs detected by RNAseq were also predicted by at least one, if not more, of the computational programs we used. We validated 14 small RNAs from different categories in our analysis using RT-PCR. We have also shown that synthetic sRNAs can be used in A. succinogenes to decrease gene expression. While our first attempt at designing and using a synthetic sRNA was successful as a proof of concept, much optimization is needed in the synthetic sRNA design and promoter strength as well. Further optimization of these synthetic sRNAs would likely entail changing the length and binding energy of the target regions or even testing several different scaffold sequences. One experiment that is likely to give us more insight into Hfq-binding sRNAs is a copurification experiment using Hfq-his6. We plan to sequence the Hfq-binding RNAs obtained from this experiment to confirm the Hfq-binding sRNAs identified in this study, maybe identify more Hfq-binding sRNAs, and identify some of the target mRNAs. Our study is a great first step towards understanding sRNAs and realizing their potential as metabolic tools for engineering A. succinogenes and studying other pasteurellaceae species. 3.5 Acknowledgements I thank Justin Beauchamp for his hands on help with the initial troubleshooting for analysis of the RNAseq. Rhiannon LeVeque for sharing the Mulk’s laboratory RNA purification protocol. 117 I thank Dr. Martha Mulks for her helpful discussions in the initial planning phase of this study. I thank Dr. Ben Johnson and John Johnson for helpful discussions regarding RNA analysis. I am grateful to Jeff Landgraf of the MSU Genomics Core for making the small RNA libraries and sequencing. 118 APPENDIX 119 Table A3.1 Manually identified smRNAs using alignment files created by Rockhopper Candidate smRNA146 smRNA147 smRNA148 smRNA149 smRNA150 smRNA151 smRNA152 smRNA153 smRNA154 smRNA155 smRNA156 smRNA157 smRNA158 smRNA159 smRNA160 smRNA161 smRNA162 smRNA163 smRNA164 smRNA165 smRNA166 smRNA167 smRNA168 smRNA169 smRNA170 smRNA171 smRNA172 Rfam annotation Transcriptio n start Transcription stop Size (nt) Coding strand Neighborhood tRNA 30,584 31,859 39,498 45,565 48,319 83,116 102,424 102,498 102,542 113,439 115,771 130,180 138,915 180,930 194,098 195,459 199,493 199,497 214,454 214,473 218,099 248,074 293,615 298,474 306,273 332,486 339,352 30,640 31,907 39,529 45,614 48,364 83,215 102,470 102,582 102,588 113,530 115,918 130,330 139,034 181,103 194,153 195,540 199,572 199,570 214,513 214,508 218,172 248,108 293,774 298,517 306,369 332,548 339,413 56 48 31 49 45 99 46 84 46 91 147 150 119 173 55 81 79 73 59 35 73 34 159 43 96 62 61 + + + + + + + + + + + + + ASUC_RS00135/ASUC_RS00140 ASUC_RS00140/ASUC_RS00145 ASUC_RS00185/ASUC_RS00190 ASUC_RS00220/ASUC_RS00225 ASUC_RS00230/ASUC_RS00235 ASUC_RS00360/ASUC_RS00365 ASUC_RS00455/ASUC_RS00460 ASUC_RS00455/ASUC_RS00460 ASUC_RS00455/ASUC_RS00460 ASUC_RS00530/ASUC_RS00535 ASUC_RS00535/ASUC_RS00540 ASUC_RS00600/ASUC_RS00605 ASUC_RS00630/ASUC_RS00635 ASUC_RS00810/ASUC_RS00815 ASUC_RS00875/ASUC_RS00880 ASUC_RS00880/ASUC_RS00885 ASUC_RS00895/ASUC_RS00900 ASUC_RS00895/ASUC_RS00900 ASUC_RS00965/ASUC_RS00970 ASUC_RS00965/ASUC_RS00970 ASUC_RS00975/ASUC_RS00980 ASUC_RS01110/ASUC_RS01115 ASUC_RS01350/ASUC_RS01355 ASUC_RS01375/ASUC_RS01380 ASUC_RS01415/ASUC_RS01420 ASUC_RS01540/ASUC_RS01545 ASUC_RS01585/ASUC_RS01590 120 Predicted by algorithms M Table A3.1 (cont’d) smRNA173 smRNA174 smRNA175 smRNA176 smRNA177 smRNA178 smRNA179 smRNA180 smRNA181 smRNA182 smRNA183 smRNA184 smRNA185 smRNA186 smRNA187 smRNA188 smRNA189 smRNA190 smRNA191 smRNA192 smRNA193 smRNA194 smRNA195 smRNA196 smRNA197 smRNA198 smRNA199 smRNA200 smRNA201 smRNA202 smRNA203 His_leader 6S RNA Parecho_CRE 420,735 429,830 436,423 442,689 445,536 445,542 449,450 553,135 606,527 635,596 662,062 663,380 691,158 701,823 718,766 731,105 744,521 785,398 831,957 837,804 845,242 866,148 868,810 936,289 967,596 978,135 979,854 1,000,431 1,052,438 1,059,515 1,064,546 420,767 429,921 436,943 442,725 445,584 445,585 449,585 553,183 606,776 635,688 662,237 663,492 691,339 701,942 718,947 731,329 744,551 785,589 832,102 838,075 845,452 866,268 868,875 936,468 967,728 978,207 980,036 1,000,539 1,052,472 1,059,731 1,064,610 32 91 520 36 48 43 135 48 249 92 175 112 181 119 181 224 30 191 145 271 210 120 65 179 132 72 182 108 34 216 64 121 + + + + + + + + + + + + + + + + + + + + ASUC_RS01910/ASUC_RS01915 ASUC_RS01955/ASUC_RS01960 ASUC_RS01975/ASUC_RS01980 ASUC_RS02010/ASUC_RS02015 ASUC_RS02030/ASUC_RS02035 ASUC_RS02030/ASUC_RS02035 ASUC_RS02055/ASUC_RS02060 ASUC_RS02590/ASUC_RS02595 ASUC_RS02850/ASUC_RS02855 ASUC_RS02995/ASUC_RS03000 ASUC_RS03105/ASUC_RS03110 ASUC_RS03115/ASUC_RS03120 ASUC_RS03270/ASUC_RS03275 ASUC_RS03320/ASUC_RS03325 ASUC_RS03405/ASUC_RS03410 ASUC_RS03475/ASUC_RS03480 ASUC_RS03550/ASUC_RS03555 ASUC_RS03755/ASUC_RS03760 ASUC_RS04000/ASUC_RS010990 ASUC_RS04020/ASUC_RS04025 ASUC_RS04055/ASUC_RS04060 ASUC_RS04165/ASUC_RS04170 ASUC_RS04180/ASUC_RS04185 ASUC_RS04530/ASUC_RS04535 ASUC_RS04675/ASUC_RS04680 ASUC_RS04730/ASUC_RS04735 ASUC_RS04735/ASUC_RS04740 ASUC_RS04845/ASUC_RS04850 ASUC_RS05065/ASUC_RS05070 ASUC_RS05085/ASUC_RS05090 ASUC_RS05105/ASUC_RS05110 S, M B, I B, I, Z, M Z Z Z Z I Table A3.1 (cont’d) smRNA204 smRNA205 smRNA206 smRNA207 smRNA208 smRNA209 smRNA210 smRNA211 smRNA212 smRNA213 smRNA214 smRNA215 smRNA216 smRNA217 smRNA218 smRNA219 smRNA220 smRNA221 smRNA222 smRNA223 smRNA224 smRNA225 smRNA226 smRNA227 smRNA228 smRNA229 smRNA230 smRNA231 smRNA232 smRNA233 smRNA234 TPP TPP riboswitch 1,101,346 1,136,235 1,148,049 1,169,658 1,178,550 1,185,698 1,192,341 1,192,343 1,226,566 1,249,629 1,254,045 1,258,601 1,259,001 1,350,512 1,356,570 1,377,923 1,378,631 1,379,749 1,412,500 1,430,292 1,438,223 1,440,698 1,498,787 1,520,162 1,543,389 1,607,502 1,625,036 1,628,813 1,644,162 1,654,562 1,665,202 1,101,403 1,136,327 1,148,185 1,169,830 1,178,645 1,185,896 1,192,408 1,192,388 1,226,635 1,249,785 1,254,220 1,258,640 1,259,063 1,350,633 1,356,750 1,377,967 1,378,725 1,379,827 1,412,583 1,430,344 1,438,385 1,440,777 1,498,840 1,520,316 1,543,436 1,607,619 1,625,112 1,628,940 1,644,218 1,654,858 1,665,386 57 92 136 172 95 198 67 45 69 156 175 39 62 121 180 44 94 78 83 52 162 79 53 154 47 117 76 127 56 296 184 122 + + + + + + + + + + + + + + + + + + + - ASUC_RS05270/ASUC_RS05275 ASUC_RS05410/ASUC_RS05415 ASUC_RS05455/ASUC_RS05460 ASUC_RS05560/ASUC_RS05565 ASUC_RS05600/ASUC_RS05605 ASUC_RS05625/ASUC_RS05630 ASUC_RS05655/ASUC_RS05660 ASUC_RS05655/ASUC_RS05660 ASUC_RS05800/ASUC_RS05805 ASUC_RS05910/ASUC_RS05915 ASUC_RS05930/ASUC_RS05935 ASUC_RS05950/ASUC_RS05955 ASUC_RS05955/ASUC_RS05960 ASUC_RS06255/ASUC_RS06260 ASUC_RS06285/ASUC_RS06290 ASUC_RS06450/ASUC_RS06455 ASUC_RS06455/ASUC_RS06460 ASUC_RS06465/ASUC_RS06470 ASUC_RS06640/ASUC_RS06645 ASUC_RS06720/ASUC_RS06725 ASUC_RS06755/ASUC_RS06760 ASUC_RS06770/ASUC_RS06775 ASUC_RS07045/ASUC_RS07050 ASUC_RS07125/ASUC_RS07130 ASUC_RS07245/ASUC_RS07250 ASUC_RS07530/ASUC_RS07535 ASUC_RS07625/ASUC_RS7630 ASUC_RS07635/ASUC_RS07640 ASUC_RS07700/ASUC_RS07705 ASUC_RS07745/ASUC_RS07750 ASUC_RS07805/ASUC_RS07810 I Z Z B, I, M Table A3.1 (cont’d) smRNA235 TPP riboswitch 1,668,651 smRNA236 1,707,974 smRNA237 1,712,810 smRNA238 1,724,670 smRNA239 1,741,704 smRNA240 1,825,558 smRNA241 1,837,348 smRNA242 MOCO_RNA_motif 1,838,604 smRNA243 1,851,093 smRNA244 1,862,282 smRNA245 1,881,063 smRNA246 1,968,657 smRNA247 1,985,767 smRNA248 2,001,685 smRNA249 2,013,568 smRNA250 2,038,722 smRNA251 2,089,369 smRNA252 2,144,262 smRNA253 2,144,312 smRNA254 2,177,961 smRNA255 2,181,502 smRNA256 2,222,240 smRNA257 2,258,917 smRNA258 2,271,792 smRNA259 2,282,160 smRNA260 2,286,315 B, BSRD; I, INFERNAL; M, Rfam; S, SIPHT; Z, RNAz 1,668,974 1,708,024 1,712,890 1,724,828 1,741,752 1,825,646 1,837,609 1,838,746 1,851,234 1,862,350 1,881,151 1,968,878 1,985,884 2,001,800 2,013,612 2,038,821 2,089,514 2,144,302 2,144,415 2,178,043 2,181,720 2,222,496 2,259,021 2,271,888 2,282,232 2,286,393 323 50 80 158 48 88 261 142 141 68 88 221 117 115 44 99 145 40 103 82 218 256 104 96 72 78 123 + + + + + + + + + + + ASUC_RS07825/ASUC_RS07830 ASUC_RS08025/ASUC_RS08030 ASUC_RS08055/ASUC_RS08060 ASUC_RS08120/ASUC_RS08125 ASUC_RS08185/ASUC_RS08190 ASUC_RS08605/ASUC_RS08610 ASUC_RS08675/ASUC_RS08680 ASUC_RS08680/ASUC_RS08685 ASUC_RS08730/ASUC_RS08735 ASUC_RS08780/ASUC_RS08785 ASUC_RS08865/ASUC_RS08870 ASUC_RS09250/ASUC_RS09255 ASUC_RS09360/ASUC_RS09365 ASUC_RS09445/ASUC_RS09450 ASUC_RS09495/ASUC_RS09500 ASUC_RS09605/ASUC_RS09610 ASUC_RS09840/ASUC_RS09845 ASUC_RS10070/ASUC_RS10075 ASUC_RS10070/ASUC_RS10075 ASUC_RS10215/ASUC_RS10220 ASUC_RS10325/ASUC_RS10240 ASUC_RS10420/ASUC_RS10425 ASUC_RS10620/ASUC_RS10625 ASUC_RS10665/ASUC_RS10670 ASUC_RS10725/ASUC_RS10730 ASUC_RS10750/ASUC_S10755 B, I, M B, I, M Z Z S Z Table A3.2 smRNAs common between SIPHT and RNAseq Candidate smRNA24 smRNA28 smRNA55 smRNA88 smRNA175 smRNA251 Rfam annotation Alpha_RBS CRISPR_DR34 - Predicted by SIPHT Transcription Transcription start stop 1,441,909 1,441,669 2,003,809 2,003,941 527,920 527,984 1,057,774 1,057,957 436,427 436,904 2,089,126 2,089,365 124 Size (nt) 465 132 64 183 477 239 Coding strand + + + + Table A3.3 smRNAs common between BSRD and RNAseq Candidate smRNA18 smRNA27 smRNA55 smRNA76 smRNA185 smRNA187 smRNA222 smRNA235 smRNA242 Rfam annotation SRP_bact GcvB Alpha_RBS Glycine riboswitch His_leader 6S RNA TPP riboswitch TPP riboswitch MOCO_RNA_motif Listed in BSRD Transcription Transcription Start Stop 1,215,630 1,215,727 1,862,334 1,862,537 527,966 528,079 806,377 806,566 691,203 691,334 718,765 718,947 1,412,495 1,412,588 1,668656 1,668,748 1,838,596 1,838,749 125 Size (nt) 98 204 114 180 132 183 94 221 154 Coding strand + + + + + + - Table A3.4 smRNAs common between INFERNAL and RNAseq Candidate Rfam annotation smRNA4 smRNA11 smRNA18 smRNA27 smRNA55 smRNA57 smRNA76 smRNA78 smRNA185 smRNA187 smRNA202 smRNA214 smRNA222 smRNA235 smRNA242 PyrR MicF SRP_bact GcvB Alpha_RBS Parecho_CRE Glycine Parecho_CRE His_leader 6S Parecho_CRE TPP TPP TPP MOCO_RNA_motif Predicted in INFERNAL Transcription Transcription Size (nt) start stop 431,716 431,819 104 981,537 981,628 92 1,215,630 1,215,730 101 1,862,334 1,862,537 204 527,966 528,079 114 599,496 599,595 100 806,377 806,566 189 869,491 869,593 103 691,203 691,334 132 718,765 718,947 183 1,059,499 1,059,618 120 1,254,038 1,254,175 138 1,412,495 1,412,588 94 1,668,656 1,668,748 93 1,838,596 1,838,749 154 126 Coding strand + + + + + + + + + + + - Table A3.5 smRNAs common between RNAz and RNAseq Candidate smRNA5 smRNA9 smRNA18 smRNA20 smRNA21 smRNA23 smRNA26 smRNA29 smRNA55 smRNA135 smRNA187 smRNA190 smRNA199 smRNA200 smRNA201 smRNA219 smRNA221 smRNA243 smRNA244 smRNA252 Rfam annotation SRP_bact Alpha_RBS 6S Predicted by RNAz Transcription Transcription start stop 517,020 51,7211 785,646 785,837 1,215,619 1,215,737 1,276,338 1,276,512 1,276,596 1,276,762 1,376,647 1,376,896 1,545,554 1,545,663 2,089,122 2,089,319 527,919 528,046 1,981,636 1,981,823 718,771 718,946 785,408 785,590 979,919 980,050 1,000,433 1,000,492 1,052,422 1,052,593 1,377,838 1,377,966 1,379,748 1,379,829 1,851,057 1,851,225 1,862,313 1,862,473 2,144,253 2,144,349 127 Size (nt) 192 192 119 175 167 250 110 198 128 188 176 183 132 60 172 129 82 169 161 97 Coding strand + + + + + + + + + + + - Table A3.6 smRNAs common between Rfam and RNAseq Candidate smRNA18 smRNA27 smRNA55 smRNA77 smRNA146 smRNA175 smRNA187 smRNA222 smRNA235 smRNA242 Rfam annotation SRP_bact GcvB Alpha_RBS Thr_leader tRNA CRISPR-DR34 6S TPP TPP MOCO_RNA_motif Predicted by BLAST against Rfam Transcription Transcription Size (nt) start stop 1,215,630 1,215,730 101 1,862,318 1,862,528 211 5,279,66 5,28,079 114 8,58,744 8,58,862 119 30,593 30,634 42 4,36,472 4,36,903 36 718,765 718,947 183 1,412,505 1,412,583 79 1,668,661 1,668,738 78 1,838,596 1,838,749 154 128 Coding strand + + + + + - Table A3.7 Primers used for RT-PCR validation of sRNAs Primer Sequence smRNA4_F smRNA4_R smRNA8_F smRNA8_R smRNA11_F smRNA11_R smRNA18_F smRNA18_R smRNA21_F smRNA21_R smRNA27_F smRNA27_R smRNA28_F smRNA28_R smRNA47_F smRNA47_R smRNA49_F smRNA49_R smRNA55_F smRNA55_R smRNA76_F smRNA76_R smRNA126_F smRNA126_R smRNA135_F smRNA135_R smRNA187_F smRNA187_R ATTGTAATGGCACTGCGAAATG AAAAAACGATGCCCCCTGAG AAGCAACTTGCAGTTGTG AAAAAAAACCGCTCTTTTCCTTG AACGTTGTGATGATGTTAAACG AAAAAAAGCCGGACTAAGGGAAT ACTAAGCCGGTGTGCGAAAG GAAACCTCCCCAGTGATTC AAAATTAACCGCACTTTATGTTC ACAAGTAAACTCGCTACGC TATTTCTAGTTTGTCCGCTCT AACACCGCACTTTTTTGGAG ATTCAAACAATAATAGATAATCACTCC AAGAAAAAGGCGTTTCCTTGTG TGCAATTCGTCGGAGATTTG ATTGAACAATCCGAAAGTTCC ATGCCGACTTTAGGACAAAAG CTTACGGCGAGGGTATTC ACAGGTTGAGCAGTTATACTG GCACTCCTATATTTTAACTAATTTG CAATAAAAAACTGTTCGGACGAAGG GGCAGAAGAGATAAAAATGATAGG GAATTTGGCAGAGAAGTAGTAC AATACCCAAGCGGTCCTC TTTAGCAAGTGCGGTCAG AAAGGCAACCCTGCTTTAC ATTACCTGAGATGCTCGCCAGC GGAGTCAGTTGTAACCGTTTTTAGG 129 Expected product size (n) 99 91 85 146 105 244 70 120 141 114 240 93 193 128 Table A3.8 Primers and g-blocks used in this study Primer Sequence Specificity (direction) CV662a CATACGGATCCCGAAAATTTGATCTAGTTAACATTTTTTAGGTATAAATAGTTTT AAAATAGATCTAGTTTGGAAGAATCATGCTGAACTCCTTAAACTTTCGGCGTTTC TCTCCCCACAAGGAAACGCCTTTTTCTTCTAGAATCGA CV663a TCGATTCTAGAAAAAAAAACCGCTCTTTTCCTTGCGGAATTGAATGATTCTGCCA ACCTACTTTGAATCCAAACTAGATCTATTTTAAAACTATTTATACCTAAAAAATG TTAACTAGATCAAATTTTCGGGATCCGTATG CV664a CATACGGATCCCGAAAATTTGATCTAGTTAACATTTTTTAGGTATAAATAGTTTT AAAATAGATCTAGTTTGGATTCAAAGTAGGTTGGCAGAATCATAATCACTCCAAC TTTCGGCGTTTCTCTCCCCACAAGGAAACGCCTTTTTCTTCTAGAATCGA a CV665 TCGATTCTAGAAGAAAAAGGCGTTTCCTTGTGGGGAGAGAAACGCCGAAAGTTGG AGTGATTATGATTCTGCCAACCTACTTTGAATCCAAACTAGATCTATTTTAAAAC TATTTATACCTAAAAAATGTTAACTAGATCAAATTTTCGGGATCCGTATG a CV666 CATACGGATCCCGAAAATTTGATCTAGTTAACATTTTTTAGGTATAAATAGTTTT AAAATAGATCTAGTTTGGATTCAAAGTAGGTTGGCAGAATCATAACTTTCGGCGT TTCTCTCCCCACA AGGAAACGCCTTTTTCTTCTAGAATCGA CV667a TCGATTCTAGAAGAAAAAGGCGTTTCCTTGTGGGGAGAGAAACGCCGAAAGTTAT GATTCTGCCAACCTACTTTGAATCCAAACTAGATCTATTTTAAAACTATTTATAC CTAAAAAATGTTAACTAGATCAAATTTTCGGGATCCGTATG CV668a CATACGGATCCCGAAAATTTGATCTAGTTAACATTTTTTAGGTATAAATAGTTTT AAAATAGATCTAGTTTGGATAGAATCATGCTGAACTCCTTAAACTTTCGGCGTTT CTCTCCCCACAAG GAAACGCCTTTTTCTTCTAGAATCGA CV669a TCGATTCTAGAAGAAAAAGGCGTTTCCTTGTGGGGAGAGAAACGCCGAAAGTTTA AGGAGTTCAGCATGATTCTATCCAAACTAGATCTATTTTAAAACTATTTATACCT AAAAAATGTTAACTAGATCAAATTTTCGGGATCCGTATG CV677 AGCGCCTGATGCGGGGATCCCGAAAATTTGATCTAGTTAACATTTTTTAG smRNA_lacZ1 (F) CV678 CCAGTGAATTCGAGCTCAAAAAAAACCGCTCTTTTC CV679 CCAGTGAATTCGAGCTCAGAAAAAGGCGTTTCCTTG CV640 CV695a,b CV696a,b CV697a,b CV698a,b a b smRNA_lacZ1 (R) smRNA_lacZ2 (F) smRNA_lacZ2 (R) smRNA_lacZ3 (F) smRNA_lacZ3 (R) smRNA_lacZ4 (F) smRNA_lacZ4 (R) smRNA_lacZ1 (F) smRNA_lacZ1 (R) smRNA_lacZ2, smRNA_lacZ3, smRNA_lacZ4 (R) CTTCGCTATTACGCCAGCTG Sequencing primer for pLGZ920 (90 bp downstream of SacI site) CATACGGATCCCGAAAATTTGATCTAGTTAACATTTTTTAGGTATAAATAGTTTT smRNA_ackA1 AAAATAGATCTAGTTTGGACAGTTAAGAATTAAAACTAATTTGGACATTCAATTC CGCAAGGAAAAGAGCGGTTTTTTTTTCTAGAATCGA CATACGGATCCCGAAAATTTGATCTAGTTAACATTTTTTAGGTATAAATAGTTTT smRNA_ackA2 AAAATAGATCTAGTTTGGACAGTTAAGAATTAAAACTAATTTGGACATAATCACT CCAACTTTCGGCGTTTCTCTCCCCACAAGGAAACGCCTTTTTCTTCTAGAATCGA CATACGGATCCCGAAAATTTGATCTAGTTAACATTTTTTAGGTATAAATAGTTTT smRNA_pta1 AAAATAGATCTAGTTTGGAGGGATAAGAATAAATGTACGAGACATTCAATTCCGC AAGGAAAAGAGCGGTTTTTTTTTCTAGAATCGA CATACGGATCCCGAAAATTTGATCTAGTTAACATTTTTTAGGTATAAATAGTTTT smRNA_pta2 AAAATAGATCTAGTTTGGAGGGATAAGAATAAATGTACGAGACATAATCACTCCA ACTTTCGGCGTTTCTCTCCCCACAAGGAAACGCCTTTTTCTTCTAGAATCGA Restriction sites BamHI and XbaI underlined Sequences for g-blocks. 130 REFERENCES 131 REFERENCES 1. Itoh T, Tomizawa J-I. 1980. Formation of an RNA primer for initiation of replication of ColE1 DNA by ribonuclease H. Proc Natl Acad Sci USA 77:2450-2454. 2. Stougaard P, Molin S, Nordström K. 1981. RNAs involved in copy-number control and incompatibility of plasmid R1. Proc Natl Acad Sci USA 78:6008-6012. 3. Gottesman S, Storz G. 2011. Bacterial small RNA regulators: versatile roles and rapidly evolving variations. Cold Spring Harb Perspect Biol 3:1-16. 4. Moller T, Franch T, Hojrup P, Keene DR, Bachinger HP, Brennan RG, ValentinHansen P. 2002. Hfq: a bacterial Sm-like protein that mediates RNA-RNA interaction. Mol Cell 9:23-30. 5. Mandal M, Boese B, Barrick JE, Winkler WC, Breaker RR. 2003. Riboswitches control fundamental biochemical pathways in Bacillus subtilis and other bacteria. Cell 113:577-586. 6. Loh E, Dussurget O, Gripenland J, Vaitkevicius K, Tiensuu T, Mandin P, Repoila F, Buchrieser C, Cossart P, Johansson J. 2009. A trans-acting riboswitch controls expression of the virulence regulator PrfA in Listeria monocytogenes. Cell 139:770-779. 7. Brouns SJJ, Jore MM, Lundgren M, Westra ER, Slijkhuis RJH, Snijders APL, Dickman MJ, Makarova KS, Koonin EV, Van Der Oost J. 2008. Small CRISPR RNAs guide antiviral defense in prokaryotes. Science 321:960-964. 8. Chao Y, Papenfort K, Reinhardt R, Sharma CM, Vogel J. 2012. An atlas of Hfq‐ bound transcripts reveals 3′ UTRs as a genomic reservoir of regulatory small RNAs. EMBO J 31:4005-4019. 9. Chao Y, Vogel J. 2016. A 3' UTR-derived small RNA provides the regulatory noncoding arm of the inner membrane stress response. Mol Cell 61:352-363. 10. Chao Y, Papenfort K, Reinhardt R, Sharma CM, Vogel J. 2012. An atlas of Hfqbound transcripts reveals 3' UTRs as a genomic reservoir of regulatory small RNAs. EMBO J 31:4005-4019. 11. Bouvier J, Stragier P, Morales V, Remy E, Gutierrez C. 2008. Lysine represses transcription of the Escherichia coli dapB gene by preventing its activation by the ArgP activator. J Bacteriol 190:5224-5229. 12. Soutourina OA, Monot M, Boudry P, Saujet L, Pichon C, Sismeiro O, Semenova E, Severinov K, Le Bouguenec C, Coppee JY, Dupuy B, Martin-Verstraete I. 2013. 132 Genome-wide identification of regulatory RNAs in the human pathogen Clostridium difficile. PLoS Genet 9:e1003493. 13. McClure R, Tjaden B, Genco C. 2014. Identification of sRNAs expressed by the human pathogen Neisseria gonorrhoeae under disparate growth conditions. Front Microbiol 5:456-456. 14. Sharma R, Arya S, Patil SD, Sharma A, Jain PK, Navani NK, Pathania R. 2014. Identification of novel regulatory small RNAs in Acinetobacter baumannii. PloS one 9:e93833. 15. Livny J, Brencic A, Lory S, Waldor MK. 2006. Identification of 17 Pseudomonas aeruginosa sRNAs and prediction of sRNA-encoding genes in 10 diverse pathogens using the bioinformatic tool sRNAPredict2. Nucleic Acids Res 34:3484-3493. 16. Schroeder CL, Narra HP, Sahni A, Rojas M, Khanipov K, Patel J, Shah R, Fofanov Y, Sahni SK. 2016. Identification and characterization of novel small RNAs in Rickettsia prowazekii. Front Microbiol 7:859. 17. Ramos CG, Grilo AM, da Costa PJ, Leitao JH. 2013. Experimental identification of small non-coding regulatory RNAs in the opportunistic human pathogen Burkholderia cenocepacia J2315. Genomics 101:139-48. 18. Mann B, van Opijnen T, Wang J, Obert C, Wang YD, Carter R, McGoldrick DJ, Ridout G, Camilli A, Tuomanen EI, Rosch JW. 2012. Control of virulence by small RNAs in Streptococcus pneumoniae. PLoS Pathog 8:e1002788. 19. Christiansen JK, Nielsen JS, Ebersbach T, Valentin-Hansen P, Sogaard-Andersen L, Kallipolitis BH. 2006. Identification of small Hfq-binding RNAs in Listeria monocytogenes. RNA 12:1383-96. 20. Arnvig KB, Young DB. 2009. Identification of small RNAs in Mycobacterium tuberculosis. Mol Microbiol 73:397-408. 21. Kuhnert P, Christensen H. 2008. Pasteurellaceae: biology, genomics and molecular aspects. Horizon Scientific Press. 22. Santana EA, Harrison A, Zhang X, Baker BD, Kelly BJ, White P, Liu Y, Munson Jr RS. 2014. HrrF is the Fur-regulated small RNA in nontypeable Haemophilus influenzae. PloS one 9:e105644. 23. Baddal B, Muzzi A, Censini S, Calogero RA, Torricelli G, Guidotti S, Taddei AR, Covacci A, Pizza M, Rappuoli R. 2015. Dual RNA-seq of nontypeable Haemophilus influenzae and host cell transcriptomes reveals novel insights into host-pathogen cross talk. mBio 6:e01765-15. 133 24. Sridhar J, Sekar K, Rafi ZA. 2009. CsrA interacting small RNAs in Haemophilus spp genomes: a theoretical analysis. Arch Microbiol 191:451-459. 25. Rossi CC, Bossé JT, Li Y, Witney AA, Gould KA, Langford PR, Bazzolli DMS. 2016. A computational strategy for the search of regulatory small RNAs in Actinobacillus pleuropneumoniae. RNA 22:1373-1385. 26. Guettler MV, Rumler D, Jain MK. 1999. Actinobacillus succinogenes sp. nov., a novel succinic-acid-producing strain from the bovine rumen. Int J Syst Bacteriol 49:207–216. 27. Cho C, Lee SY. 2016. Efficient gene knockdown in Clostridium acetobutylicum by synthetic small regulatory RNAs. Biotechnol Bioeng 114:374-383. 28. Na D, Yoo SM, Chung H, Park H, Park JH, Lee SY. 2013. Metabolic engineering of Escherichia coli using synthetic small regulatory RNAs. Nat Biotechnol 31:170-174. 29. McKinlay JB, Zeikus JG, Vieille C. 2005. Insights into Actinobacillus succinogenes fermentative metabolism in a chemically defined growth medium. Appl Environ Microbiol 71:6651-6656. 30. Schindler BD, Joshi RJ, Vieille C. 2014. Respiratory glycerol metabolism of Actinobacillus succinogenes 130Z for succinate production. J Ind Microbiol Biotechnol 41:1339-1352. 31. Kim P, Laivenieks M, McKinlay J, Vieille C, Zeikus JG. 2004. Construction of a shuttle vector for the overexpression of recombinant proteins in Actinobacillus succinogenes. Plasmid 51:108–115. 32. Bolger AM, Lohse M, Usadel B. 2014. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics 30:2114-20. 33. Tjaden B. 2015. De novo assembly of bacterial transcriptomes from RNA-seq data. Genome biology 16:1. 34. McClure R, Balasubramanian D, Sun Y, Bobrovskyy M, Sumby P, Genco CA, Vanderpool CK, Tjaden B. 2013. Computational analysis of bacterial RNA-Seq data. Nucleic Acids Res 41:e140-e140. 35. Busch A, Richter AS, Backofen R. 2008. IntaRNA: efficient prediction of bacterial sRNA targets incorporating target site accessibility and seed regions. Bioinformatics 24:2849-2856. 36. Wright PR, Georg J, Mann M, Sorescu DA, Richter AS, Lott S, Kleinkauf R, Hess WR, Backofen R. 2014. CopraRNA and IntaRNA: predicting small RNA targets, networks and interaction domains. Nucleic Acids Res 42:W119-23. 134 37. Gruber AR, Lorenz R, Bernhart SH, Neuböck R, Hofacker IL. 2008. The Vienna RNA Websuite. Nucleic Acids Res 36:W70-W74. 38. Livny J, Teonadi H, Livny M, Waldor MK. 2008. High-throughput, kingdom-wide prediction and annotation of bacterial non-coding RNAs. PLoS One 3:e3197. 39. Li L, Huang D, Cheung MK, Nong W, Huang Q, Kwan HS. 2013. BSRD: a repository for bacterial small regulatory RNA. Nucleic Acids Res 41:D233-D238. 40. Nawrocki EP, Eddy SR. 2013. Infernal 1.1: 100-fold faster RNA homology searches. Bioinformatics 29:2933-2935. 41. Gruber AR, Findeiss S, Washietl S, Hofacker IL, Stadler PF. 2010. RNAz 2.0: improved noncoding RNA detection. Pac Symp Biocomput:69-79. 42. Nawrocki EP, Burge SW, Bateman A, Daub J, Eberhardt RY, Eddy SR, Floden EW, Gardner PP, Jones TA, Tate J, Finn RD. 2015. Rfam 12.0: updates to the RNA families database. Nucleic Acids Res 43:D130-D137. 43. Cros MJ, de Monte A, Mariette J, Bardou P, Grenier-Boley B, Gautheret D, Touzet H, Gaspin C. 2011. RNAspace.org: An integrated environment for the prediction, annotation, and analysis of ncRNA. RNA 17:1947-1956. 44. Grissa I, Vergnaud G, Pourcel C. 2007. CRISPRFinder: a web tool to identify clustered regularly interspaced short palindromic repeats. Nucleic Acids Res 35:W52-W57. 45. McKinlay JB, Shachar-Hill Y, Zeikus JG, Vieille C. 2007. Determining Actinobacillus succinogenes metabolic pathways and fluxes by NMR and GC-MS analyses of 13Clabeled metabolic product isotopomers. Metab Eng 9:177–192. 46. McKinlay JB, Laivenieks M, Schindler BD, Mckinlay AA, Siddaramappa S, Challacombe JF, Lowry SR, Clum A, Lapidus AL, Burkhart KB, Harkins V, Vieille C. 2010. A genomic perspective on the potential of Actinobacillus succinogenes for industrial succinate production. BMC Genomics 11:680. 47. Hofacker IL, Fontana W, Stadler PF, Bonhoeffer LS, Tacker M, P. S. 1994. Fast folding and comparison of RNA secondary structures. Monatshefte für Chemie/Chemical Monthly 125:167-188. 48. Pulvermacher SC, Stauffer LT, Stauffer GV. 2009. Role of the Escherichia coli Hfq protein in GcvB regulation of oppA and dppA mRNAs. Microbiology 155:115-123. 49. Ishikawa H, Otaka H, Maki K, Morita T, Aiba H. 2012. The functional Hfq-binding module of bacterial sRNAs consists of a double or single hairpin preceded by a U-rich sequence and followed by a 3' poly(U) tail. RNA 18:1062-1074. 135 50. Otaka H, Ishikawa H, Morita T, Aiba H. 2011. PolyU tail of rho-independent terminator of bacterial small RNAs is essential for Hfq action. Proc Natl Acad Sci U S A 108:13059-13064. 51. Udekwu KI, Darfeuille F, Vogel J, Reimegard J, Holmqvist E, Wagner EG. 2005. Hfq-dependent regulation of OmpA synthesis is mediated by an antisense RNA. Genes Dev 19:2355-2366. 52. Yoo SM, Na D, Lee SY. 2013. Design and use of synthetic regulatory small RNAs to control gene expression in Escherichia coli. Nat Protocols 8:1694-1707. 53. van der Werf MJ, Guettler MV, Jain MK, Zeikus JG. 1997. Environmental and physiological factors affecting the succinate product ratio during carbohydrate fermentation by Actinobacillus sp. 130Z. Arch Microbiol 167:332–342. 136 Chapter 4 Development of a markerless knockout method for Actinobacillus succinogenes 137 This work is previously published in Applied and Environmental Microbiology as: Joshi RV, Schindler BD, McPherson N, Tiwari K, Vieille C (2014) Development of a Markerless Knockout Method for Actinobacillus succinogenes. Appl Environ Microbiol 80:3053–3061. doi:10.1128/AEM.00492-14 The article can be viewed at the following URL: http://aem.asm.org/content/80/10/3053 138 The work described in this chapter was partly completed as a requirement for a professional master’s degree and was submitted as a form of a report titled, “Development of a gene knockout method for A. succinogenes strain 130Z and engineering of A. succinogenes for aspartate production”. My contributions to this paper during my PhD are listed as below:  Construction of double knockouts ΔpflB ΔlacZ::icd, ΔpflB Δcit::icd, ΔpflB Δacn::icd.  Training and mentoring Kanupriya Tiwari during her research project in the summer of 2013. She worked on determining the effects of the DNA construct length and of the length of the homologous regions on the efficiency of homologous recombination in A. succinogenes.  Drafting parts, reading, and editing manuscript. 139 4.1 Abstract Actinobacillus succinogenes is one of the best natural succinate-producing organisms, but it still needs engineering to further increase succinate yield and productivity. In this study we developed a markerless knockout method for A. succinogenes using natural transformation or electroporation. The Escherichia coli isocitrate dehydrogenase gene with flanking flippase recognition target sites was used as the positive selection marker, making use of A. succinogenes’s auxotrophy for glutamate to select for growth on isocitrate. The Saccharomyces cerevisiae flippase recombinase (Flp) was used to remove the selection marker, allowing its reuse. Finally, the plasmid expressing flp was cured using acridine orange. We demonstrate that at least two consecutive deletions can be introduced into the same strain using this approach, that no more than a total of 1 kb of DNA is needed on each side of the selection cassette to protect from endonucleases activity during transformation, and that no more than 200 bp of homologous DNA is needed on each side for efficient recombination. We also demonstrate that electroporation can be used as an alternative transformation method to obtain knockout mutants, and that an enriched defined medium can be used for direct selection of knockout mutants on agar plates with high efficiency. Knockout mutants of the fumarate reductase and the citrate lyase operons, as well as of the pyruvate formate lyase, β-galactosidase, and aconitase genes were obtained using this knockout strategy. 140 4.2 Introduction Dicarboxylic acids are among the top of the US’s Department of Energy list of valueadded chemicals from biomass (1). In particular, if produced economically by fermentation, succinic acid could replace maleic anhydride as the precursor to many bulk and commodity chemicals, with a potential market of 25 billion tons per year (2). Actinobacillus succinogenes strain 130Z (ATCC 55618) is a facultative anaerobe and a member of the Pasteurellaceae. It is the highest natural succinate producer known (3-5). A. succinogenes produces succinate, acetate, and formate as its major fermentative products with ethanol as a minor by-product (6). Under optimized conditions, wild-type A. succinogenes produces 80 g L-1 succinate, while chemically induced mutant strains resistant to fluoroacetate produce up to 110 g L-1 succinate (3), suggesting that even higher succinate yields can be obtained with genetically engineered strains. Bio-based succinate will only be price-competitive with petroleum-based maleic anhydride if alternative fermentation products can be eliminated. Achieving a homosuccinate fermentation would drastically reduce the cost of downstream succinate purification (7). Until recently, genetic tools for A. succinogenes were limited to the expression vector pLGZ920 and electroporation (8). Plasmid pLGZ920 confers ampicillin (Amp) resistance, replicates in A. succinogenes and Escherichia coli, and allows high-level expression of foreign genes in A. succinogenes from the strong, constitutive A. succinogenes pckA promoter (ppckA) (8). Electroporation of A. succinogenes with pLGZ920 yields transformants with an efficiency of 104‒106 CFU/μg of plasmid, depending on the electroporation parameters (8) (Schindler and Vieille, unpublished results). Early attempts to construct knockout mutants of A. succinogenes by allelic exchange used electroporation with suicide vectors containing gene knockout constructs interrupted by antibiotic resistance genes. The high frequency of spontaneous antibiotic resistant 141 mutants masked the low frequency of double recombination events, and no knockout mutants were isolated in these studies (McKinlay and Vieille, unpublished results). Another selection method and possibly other means to introduce DNA into cells are thus needed to develop a knockout method for A. succinogenes. In other Pasteurellaceae species conjugation and natural transformation have been commonly used to construct knockout mutants. Conjugation has been used in A. pleuropneumoniae (9), Mannheimia haemolytica, Pasteurella multocida, and Haemophilus somnus (10). Deletions were typically selected for with an antibiotic resistance marker, and SacB-based sucrose counterselection was used to select for double recombination. The conjugated plasmids were either suicide vectors or temperature-sensitive shuttle plasmids (10). One study used the Cre-lox system to remove the selection marker (11). Natural transformation has been used in many studies of H. influenzae and other naturally competent Pasteurellaceae. Natural competence is induced by starvation stress. Transformation frequencies can be as high as 10-2 (12). Natural transformation is well-suited for strain engineering since it works best with linear DNA (e.g., PCR products), which requires double recombination events for complete chromosome integration. Naturally competent Pasteurellaceae preferentially take up DNA from their own species over unrelated DNA (13, 14). They do so with a membrane-bound DNA uptake machinery that specifically recognizes and binds a conserved uptake signal sequence (USS) (15). Low transformation efficiencies of A. actinomycetemcomitans were observed with DNA fragments not containing a USS (16). All Pasteurellaceae genomes sequenced contain USS repeats, in numbers ranging from 41 to 1,760 (1,690 in A. succinogenes) (17), even though not all these species are naturally competent (12). A. succinogenes’s genome contains twenty three of the 142 twenty five genes in H. influenzae’s natural competence regulon (17, 18), including those encoding the competence regulatory proteins Sxy and CRP (cyclic AMP receptor protein). The missing two genes encode hypothetical proteins with unknown roles in natural competence (12). The frequency of USS repeats in A. succinogenes’s genome and the likely presence of a complete natural competence machinery suggest that A. succinogenes is naturally competent. In this study, we demonstrate that natural competence can be used to create knockout mutants of A. succinogenes, that the E. coli isocitrate dehydrogenase gene can be used as a positive selection marker, that the Saccharomyces cerevisiae Flp recombinase (encoded by flp) (19) can be used in A. succinogenes to remove the positive selection marker flanked by two direct Flp recognition target sites (FRT) (20), and that more than one deletion can be introduced in the same strain. We also demonstrate that no more than a total of 1 kb of DNA is needed on each side of the selection cassette to protect from exonucleases activity during transformation, and that no more than 200 bp of homologous DNA is needed for efficient recombination. The genes encoding fumarate reductase and pyruvate formate lyase were chosen as initial targets for knockout constructs. Fumarate reductase (encoded by frdABCD) converts fumarate to succinate. Knocking out fumarate reductase would likely be the first step needed to engineer A. succinogenes into a fumarate or aspartate producer. During A. succinogenes’s fermentative growth on glucose, pyruvate formate lyase (PFL, encoded by pflB) is the main enzyme converting pyruvate into acetyl-CoA, with the concomitant production of formate (21, 22). Acetyl-CoA is then the precursor of acetate and ethanol (23). Fluoroacetate-resistant mutants devoid of PFL activity are not affected in their growth and produce increased amounts of succinate (3). A pflB knockout mutation is likely to be the first mutation needed to engineer a homosuccinate-producing A. succinogenes strain (24-27). Three other sets of genes (encoding β- 143 galactosidase, citrate lyase, and aconitase) deemed non-essential for fermentative growth on glucose were targeted for deletion as well. 4.3 Materials and Methods 4.3.1 Strains, media, culture conditions, and chemicals. Strains used in this study are listed in Table 4.1. E. coli strains were cultivated in lysogeny broth (LB) and on LB agar plates (28). A. succinogenes strains were cultivated in medium B (in g L-1: yeast extract, 5; bactotryptone, 10; NaH2PO4·H2O, 8.5; K2HPO4, 15.5; NaHCO3, 2.1; and glucose, 9), AM3 defined medium (29), or Bacto brain heart infusion (BHI; Becton-Dickinson and Co., Franklin Lakes, NJ). AM3 is a phosphate-based medium containing glucose, NH4+, ten vitamins, and the amino acids glutamate, cysteine, and methionine. After natural transformations, A. succinogenes was grown in liquid AM2 (AM3 minus glutamate) in the absence or presence of isocitrate (30 mM; AM2-isocitrate) or on AM16-isocitrate or AM17 agar plates. AM17 is modified AM3 containing twice the amount of vitamins, cysteine, and methionine normally present in AM3, plus an amino acid supplement mix (in mg L-1, final concentrations in AM17: alanine, 120; aspartate, 105; asparagine·H2O, 120; glycine, 115; histidine, 21; isoleucine, 58; leucine, 90; lysine(HCl)2, 113; phenylalanine, 45; serine, 44; threonine, 56; tryptophan, 13; tyrosine, 28; and valine, 82). AM16 is AM17 without glutamate. A. succinogenes liquid cultures were grown in N2-flushed 28-mL anaerobic tubes at 37°C, with shaking at 250 rpm. OD660 of A. succinogenes cultures was monitored using a Spectronic 20 (Bausch and Lomb, Rochester, NY). To isolate single colonies, A. succinogenes strains were grown on LB plates containing 10 g L-1 glucose (27) and 10 μg mL-1 kanamycin (Km, A. succinogenes is naturally resistant to Km). Amp was added to all growth media for plasmid maintenance when required (40 μg mL-1 for A. 144 succinogenes and 100 μg mL-1 for E. coli). All cultures of A. succinogenes on agar plates were grown under a CO2-enriched atmosphere. Unless stated otherwise, chemicals were from SigmaAldrich (St. Louis, MO). Table 4.1 Strains and plasmids used in this study Description Source E. coli K-12 (ATCC19020) Wild-type strain Laboratory collection DH5α F- φ80lacZΔM15 Δ(lacZYA-argF) U169 recA1 endA1 hsdR17 Laboratory collection TOP10 (rk-, mk+) phoA supE44 λ- thi-1 gyrA96 relA1 F- mcrA Δ(mrrhsdRMS-mcrBC) φ80lacZΔM15 ΔlacΧ74 recA1 araD139 Δ(araleu) 7697 galU galK rpsL (StrR) endA1 nupG λ- Invitrogen A. succinogenes 130Z (ATCC55618) Wild-type strain ATCC Δfrd::icd 130Z derivative, contains the AscI-FRT-ppckA-icd-FRT-AscI cassette This study in the frdAB deletion Δfrd 130Z derivative, frdAB deletion, contains one FRT site ΔpflB::icd 130Z derivative, contains the AscI-FRT-ppckA-icd-FRT-AscI cassette This study in the pflB deletion ΔpflB ΔpflB-ΔlacZ::icd ΔpflB-ΔcitDEF ΔpflB-Δacn 130Z derivative, pflB deletion, contains one FRT site ΔpflB derivative, lacZ deletion, still contains the icd cassette ΔpflB derivative, citDEF deletion ΔpflB derivative, acn deletion This study This study This study This study pCR2.1 AmpR, KmR, lacZα, cloning vector Invitrogen pLGZ920 E. coli-A. succinogenes shuttle vector; AmpR; A. succinogenes ppckA (8) pCP20 pSC101 derivative, flp+, λ cI857+, λ pR Repts, AmpR, CmR (36) pCR2.1-icd pCR2.1 derivative, E. coli icd under control of A. succinogenes ppckA, flanked by FRT repeats and AscI restriction sites This study pCR2.1-Δfrd::icd pCR2.1 derivative, A. succinogenes Δfrd::FRT-ppckA-icd-FRT This study pCR2.1-Δfrd pCR2.1 derivative, frdAB deletion (2.1-kb fusion product of frdup and frdCD) This study This study Plasmids 145 Table 4.1 (cont’d) pCR2.1-ΔpflB pCR2.1 derivative, pflB deletion (2-kb fusion product of pflBup and This study pflBdown) pCR2.1-ΔpflB::icd pCR2.1 derivative, A. succinogenes ΔpflB::FRT-ppckA-icd-FRT This study pLGZ924 pLGZ920 derivative, E. coli icd under control of A. succinogenes ppckA promoter This study pCV933 pLGZ920 derivative, S. cerevisiae flp under control of A. succinogenes ppckA pCR2.1 derivative, lacZ deletion (fusion product of lacZup and lacZdown) pCR2.1 derivative, A. succinogenes ΔlacZ::FRT-ppckA-icd-FRT pCR2.1 derivative, citDEF deletion (fusion product of citDEFup and citDEFdown) pCR2.1 derivative, A. succinogenes ΔcitDEF::FRT-ppckA-icd-FRT pCR2.1 derivative, lacZ deletion (fusion product of acnup and acndown) pCR2.1 derivative, A. succinogenes Δacn::FRT-ppckA-icd-FRT This study pCR2.1-ΔlacZ pCR2.1-ΔlacZ::icd pCR2.1-Δcit pCR2.1-Δcit::icd pCR2.1-Δacn pCR2.1-Δacn::icd This study This study This study This study This study This study 4.3.2 Plasmids, DNA manipulations, and electroporations. Plasmids used in this study are listed in Table 4.1. PCR products were cloned into pCR2.1 using the TOPO-TA cloning kit (Invitrogen, Carlsbad, CA). pLGZ920 was used to express foreign genes in A. succinogenes under control of ppckA (8). One Shot TOP10 chemically competent E. coli cells (Invitrogen) were transformed as described by the manufacturer. Electrocompetent A. succinogenes cells used for transformation of linear DNA were prepared as described (28). Electrocompetent A. succinogenes cells used for transformation of circular DNA were prepared using a method modified from (30) as follows. A 10-mL culture of actively growing A. succinogenes (OD660 0.5 ± 0.2) in medium B or AM3 was incubated on ice for 10 min. Six mL of culture were harvested by centrifugation (3 min, 4,500 × g) at 4°C. The pellet was washed three times with 1 mL of cold 272 mM sucrose and finally resuspended in 100 μL of 272 mM sucrose. Electroporation was performed with 1 μL DNA (100−300 ng for circular DNA, 146 700−800 ng for linear DNA) and 100 μL electrocompetent cells in a 2-mm gap width electroporation cuvette (Bio-rad, Hercules, CA). After electroporation (settings: 25 μF, 400 Ω for plasmids, 600 Ω for linear DNA, and 2.5 kV on a Bio-Rad GenePulser™), electroporation mixtures were incubated with 0.5 mL of super optimal broth with catabolite repression (SOC) outgrowth medium (New England Biolabs, Ipswich, MA) for 1 h at 37°C. The cells were then centrifuged (3 min, 4,500 × g), resuspended in 100 μL supernatant, and spread on a single plate. DNA manipulations used standard protocols (28). PCR reactions were performed with the Advantage HD polymerase kit (Clontech, Mountain View, CA) unless otherwise stated. Restriction enzymes were from New England Biolabs. Genomic DNA extractions were performed using the Wizard genomic DNA purification kit, plasmid DNA was purified using the Wizard SV miniprep kit, and DNA fragments were recovered from PCR mixtures and agarose gels using the Wizard SV gel and PCR Clean-Up System (Wizard kits from Promega, Madison, WI). Primers used in this work are listed in Table S4.1. Primers were synthesized by the Michigan State University Research Technology Support Facility (MSU RTSF) or by Integrated DNA Technologies (Coralville, IA). PCR and cloning accuracy were confirmed by DNA sequencing performed by the MSU RTSF or by GENEWIZ, Inc. (South Plainfield, NJ). Colony PCR was performed with Taq polymerase to confirm cloning steps. 4.3.3 Construction of plasmid pLGZ924. The isocitrate dehydrogenase gene (icd) was amplified from E. coli K-12 genomic DNA using primers P1 and P2. The PCR product was cloned into pCR2.1, and then subcloned into the XbaI and SacI sites of pLGZ920, downstream of ppckA, yielding plasmid pLGZ924. 147 4.3.4 Construction of Δfrd::icd and ΔpflB::icd mutants. The E. coli icd gene under control of ppckA in pLGZ924 was amplified with primers P3 and P4. The primers added S. cerevisiae FRT sequences (5’GAAGTTCCTATTCCGAAGTTCCTATTCTCTAGAAAGTATAGGAACTTC-3’) (20) in the same orientation plus AscI restriction sites on each side of the ppckA-icd cassette to yield a 1.6 kb, AscI-FRT-ppckA-icd-FRT-AscI, cassette (Figure 4.1). After the PCR reaction, the 100-μL mixture was heated at 95°C for 20 min. 3’ A-overhangs were added to the PCR product using Taq polymerase as described in the TOPO-TA cloning kit, before cloning into pCR2.1 to yield pCR2.1-icd. A ppckA A icd FRT FRT AscI-FRT-ppckA-icd-FRT-AscI cassette (A) A frdup 1 2 3 4 Δfrd in pCR2.1 frdCD A frdup (B) M 5.0 kb 3.8 kb ppckA A icd FRT FRT frdCD Δfrd::icd in pCR2.1 and Δfrd::icd strain 2.1 kb 3.8 kb frdA frdup frdCD frdB frdABCD region in 130Z 5.0 kb frdup FRT frdCD Δfrd strain 2.1 kb Figure 4.1 Construction of the icd selection cassette and construction of strain ∆frd. (A) Physical maps of the intermediary constructs and strains involved. A: AscI; Black triangles: primers P9 and P10 used to amplify the PCR products shown in (B). (B) PCR verification of intermediate and final constructs using primers P9-P10. PCR product amplified from the 5.7-kb Δfrd::icd PCR fragment used for natural transformation (lane 1), from A. succinogenes 130Z (lane 2), from A. succinogenes ∆frdAB::icd (lane 3), and from A. succinogenes ∆frd (lane 4). Lane M: DNA molecular markers. A Δfrd DNA construct was assembled in two steps. First, the 1.1-kb region upstream of frdABCD (frdup) was amplified with primers P5 and P6, and a 1-kb fragment containing the frdCD genes was amplified with primers P7 and P8. Primers P6 and P7 overlapped by 19 nt, 148 including an AscI site. Second, the 1.1- and 1.0-kb purified fragments were fused by PCR using nested primers P9 and P10. The 2.1-kb frdup-AscI-frdCD product (Figure 4.1) was cloned into pCR2.1, yielding pCR2.1-Δfrd. The AscI-FRT-ppckA-icd-FRT-AscI fragment from pCR2.1-icd was cloned into the AscI site of pCR2.1-Δfrd, yielding pCR2.1-Δfrd::icd (Figure 4.1). This plasmid was used as a template with primers P11 and P12 to amplify a 5.4-kb fragment that was naturally transformed into A. succinogenes 130Z to create strain Δfrd::icd. A ∆pflB construct was assembled in pCR2.1 in two steps as well. Two 1.4-kb DNA fragments upstream (pflBup) and downstream (pflBdown) of A. succinogenes pflB were amplified separately using primer pairs P13-P14 and P15-P16, respectively. Primers P14 and P15 overlapped by 21 nt, including an AscI site. The pflBup and pflBdown purified PCR fragments were then fused by PCR using nested primers P17 and P18. The 2.1-kb pflBup-AscI-pflBdown product was cloned into pCR2.1, yielding plasmid pCR2.1-ΔpflB. Finally, the AscI-FRT-ppckA-icd-FRTAscI cassette from pCR2.1-icd was cloned into the AscI site of pCR2.1-ΔpflB, yielding pCR2.1ΔpflB::icd. After restriction digest with XmnI, the 7.6-kb linear plasmid DNA was naturally transformed into A. succinogenes to create strain ΔpflB::icd. 4.3.5 Construction of the ∆pflB ΔlacZ::icd, ∆pflB Δcit::icd, and ∆pflB Δacn::icd double mutants. Strain ΔpflB was used as the host to delete lacZ, the genes encoding citrate lyase (citDEF), and the aconitase gene (acn). The strategy used was identical to that used to build Δfrd and ΔpflB. Construction of plasmids pCR2.1-ΔlacZ::icd, pCR2.1-Δcit::icd, and pCR2.1-Δacn::icd used primers P35 to P40 (ΔlacZ), p41 to p46 (Δcit), and P47 to P52 (Δacn) (Table S4.1). Δcit was flanked by 0.5 kb of its upstream and downstream regions in pCR2.1-Δcit::icd. Δacn was 149 flanked by 0.4 kb and 0.45 kb of its upstream and downstream regions, respectively, in pCR2.1Δacn::icd. ΔlacZ was flanked by 0.75 kb and 0.85 kb of its upstream and downstream regions, respectively, in pCR2.1-ΔlacZ::icd. These plasmids were used as the templates with primer pairs P11-P12 (add 1 kb of vector DNA on each side of the regions flanking the deletion) and P53-P54 (add 0.1 kb of vector DNA on each side of the regions flanking the deletion) to amplify the fragments used to naturally transform strain ΔpflB. Double recombinations in strain ΔpflB were screened by colony PCR using primer pairs P39-P40, P39-P26, and P25-P40 (ΔlacZ::icd); P45P46, P45-P25, and P26-P46 (Δcit::icd), and P51-P52, P51-P26, and P25-P52 (Δacn::icd). ΔpflB ΔlacZ::icd double mutant candidates were further confirmed by patching the colonies on LB glucose plates containing 40 μL 5-bromo-4-chloro-3-indolyl-β-D-galactopyranoside (X-gal). 4.3.6 Natural transformation. The natural transformation protocol for A. succinogenes closely resembles those for H. influenzae (31) and A. pleuropneumoniae (32). To prepare competent cells, 400 μL of an overnight culture of A. succinogenes in BHI were transferred to 35 mL fresh BHI and grown to an OD600 of 0.2–0.25. Cells were washed twice with AM3 phosphate buffer (per liter: NaH2PO4·H2O, 8.5 g; K2HPO4, 15.5 g), re-suspended in 10 mL of anaerobic MIV medium (31) containing 2 mM cAMP, and incubated at 37°C for 100 min with shaking at 100 rpm to induce competence. For natural transformation, a mixture of 1 μg DNA and 1 mL competent cells was incubated in a 37°C water bath for 25 min. A negative control without DNA was included. Two volumes of BHI were added to the transformation mixtures and incubated at 37°C with shaking (250 rpm) for 100 min. Cells were then harvested (4,500 × g, 4°C, 15 min), washed twice with 0.75 mL AM3 phosphate buffer, and resuspended in 0.75 mL AM3 phosphate buffer. Finally, 150 0.25 mL of cell suspension was used to inoculate one tube each of AM3 (positive control), AM2 (negative control), and AM2-isocitrate (selective medium). Tubes were incubated at 37°C with shaking until growth was observed in AM2-isocitrate. AM2-isocitrate cultures (0.25 mL) were streaked onto LB-glucose-Km plates to isolate putative recombinants. A method for plate selection is evaluated in the Results section. 4.3.7 Removal of the icd marker. The S. cerevisiae flp gene was cloned into pLGZ920 for expression in A. succinogenes. The flp gene was amplified from plasmid pCP20 using primers P19 and P20. The purified PCR product was cloned into the XbaI and SacI sites of pLGZ920 using the In-Fusion cloning system (Clontech). In the resulting plasmid, pCV933, flp is expressed constitutively under control of A. succinogenes ppckA. pCV933 was purified and electroporated into A. succinogenes strains Δfrd::icd, ΔpflB::icd, ∆pflB Δcit::icd, and ∆pflB Δacn::icd. Transformants were screened for excision of the icd marker by colony PCR using several primer pairs. Excision of icd yielded strains ΔpflB(pCV933), Δfrd(pCV933), ∆pflB Δcit(pCV933), and ∆pflB Δacn(pCV933), each of which contains a single FRT scar in the chromosome. The deletions and the presence of the 48-nt FRT sequence framed by two AscI sites were confirmed by sequencing these regions in the genomes of strains Δfrd and ΔpflB. 4.3.8 Curing of plasmid pCV933. An overnight culture of A. succinogenes Δfrd(pCV933) or ΔpflB(pCV933) in medium B was used to inoculate a series of medium B tubes containing 0, 10, 50, 100, and 200 μg mL-1 acridine orange (AO). After 6.5 h of growth (37°C), cultures from each tube were streaked onto 151 LB-glucose-Km plates. The resulting colonies were patched onto LB-glucose-Amp plates to screen for the loss of pCV933. Amp-sensitive (Amps) colonies were screened for plasmid curing by colony PCR using primers P21 and P22, specific for sequences upstream and downstream of the pLGZ920 multiple cloning site. 4.3.9 Enzyme assays. Enzyme assays were performed with cell extracts prepared from exponential phase cultures (OD660 between 0.6 and 1.0) grown in AM3. Cells from 10-mL cultures were harvested by centrifugation (4,500 × g, 15 min, 4˚C) and washed three times with 50 mM Tris-HCl (pH 7.4) at room temperature. For ICD and citrate lyase assays, pellets were resuspended in 0.5 mL of 50 mM Tris-HCl (pH 7.4). For fumarate reductase assays pellets were resuspended in 1 mL of 50 mM potassium phosphate (pH 7.0) containing 0.1 mM dithiothreitol. Cells were sonicated with a Branson Sonifier 450 (Danbury, CT; 50% duty cycle, level 2, 20 sec, 4 times). The lysate was centrifuged and the supernatant used for the assays as crude cell extract. Total proteins were quantified using the Bio-Rad (Hercules, CA) Protein Assay Dye Reagent Concentrate with bovine serum albumin as the standard. One unit of enzyme activity was defined as the amount of enzyme needed to convert 1.0 μmol of substrate into product per min. ICD and citrate lyase activities were assayed using a Cary 300 UV/vis spectrophotometer (Varian Instruments, Walnut Creek, CA) equipped with a Peltier system. ICD activity was assayed by monitoring NADP+ reduction at 340 nm in a 1-mL reaction mixture containing 50 mM Tris-HCl (pH 7.4), 5 mM DL-isocitric acid trisodium salt, 5 mM MgCl2, 1 mM dithiothreitol, 0.3 mM NADP+, and 90 mM NaCl. The reaction was started by adding 50 μL crude cell extract. The molar extinction coefficient of NADPH was 6,200 M-1cm-1. Fumarate reductase activity was assayed in 96-well 152 plates in a Sunrise microplate reader (Tecan, Durham, NC) in an anaerobic glove bag. The 200μL reaction contained 50 mM potassium phosphate (pH 7.0), 5 mM benzyl viologen, 0.4 mM sodium dithionite, and 5 mM sodium fumarate. The reaction was started by adding 10 μL cell extract. The activity was measured at 595 nm as the difference between the slopes before and after adding the cell extract. Measurements were taken every 30 sec with shaking between measurements. All enzyme assays were performed at 37˚C on three biological replicates. 4.3.10 Analysis of fermentation products. Overnight cultures of strains 130Z and ΔpflB in AM3-25 mM NaHCO3 were inoculated (0.25 mL) into 10 mL of fresh AM3-25 mM NaHCO3. Fermentation products in culture supernatants were quantified by high-performance liquid chromatography (HPLC) (Waters, Milford, MA) using a 300 × 7.8 mm Aminex HPX-87H column (Bio-Rad, Hercules, CA) at 30°C with 4 mM H2SO4 as the mobile phase (0.6 mL/min flow rate). Organic acids were detected with a Waters 2478 UV detector at 210 nm or 254 nm. 4.4 Results 4.4.1 A positive selection method for recombination events in A. succinogenes. To avoid using antibiotic resistance genes as selection markers, a positive selection method was developed based on A. succinogenes’s metabolism. Because it is missing citrate synthase and ICD in the citric acid cycle, A. succinogenes is auxotrophic for glutamate (29). A. succinogenes has glutamate dehydrogenase activity and is able to grow on α-ketoglutarate instead of glutamate (29). We proposed that A. succinogenes expressing ICD would be able to grow in minimal medium with isocitrate substituted for glutamate. 153 A. succinogenes 130Z(pLGZ924) could grow in AM2-isocitrate, whereas 130Z could not (data not shown). Strain 130Z(pLGZ924) grew slower in AM3 and AM2-isocitrate, (generation times of 2.08 ± 0.03 h and 5.7 ± 0.1 h, respectively) than 130Z did in AM3 (generation time of 1.66 ± 0.01 h). AM3 contains 1.4 mM glutamate. Even with 5.7 mM L-isocitrate (4-fold more than glutamate in AM3), 130Z(pLGZ924) grew in AM2-isocitrate to a maximum OD660 40% lower than in AM3. In contrast to glutamate, which can only be a precursor of biomass in 130Z, isocitrate can also be metabolized to oxaloacetate and acetyl-CoA through aconitase and citrate lyase, and it seems to be partially diverted to acetate, explaining why 130Z(pLGZ924) grows slower and to a lower final OD in AM2-isocitrate than in AM3. While the icd gene looks to be a promising positive selection marker for A. succinogenes, its introduction in A. succinogenes increases the complexity of the central metabolism, suggesting that the icd gene should be excised from knockout mutant strains. 4.4.2 Natural transformation of A. succinogenes using a Δfrd::icd construct. Two USSs were present in the Δfrd::icd construct, 650 nt upstream (in frdup) and 900 nt downstream (in frdCD) of the FRT-ppckA-icd-FRT cassette, respectively (Figure 4.1). AscI restriction sites were created on both ends of the FRT-ppckA-icd-FRT cassette and in the center of the Δfrd construct for easy insertion of the cassette into the Δfrd construct. The A. succinogenes chromosome contains only 33 AscI sites, allowing our selection cassette to be used for almost any gene deletion in the genome. DNA starts being degraded by exonucleases as soon as it enters the cell. To protect the A. succinogenes sequences (~1 kb on each side of the selection marker) from degradation and maximize the chances of recombination, the PCR product used to construct strain Δfrd::icd contained 1 kb of pCR2.1 DNA on each side of the knockout construct. 154 To determine whether and when cAMP is needed to induce A. succinogenes competence, different growth and competence induction conditions were tested. cAMP was completely omitted in the first experiment. In the second experiment, 2 mM cAMP was added to BHI only. In the third experiment, 2 mM cAMP was added to MIV only. In each of these experiments the transformation mixtures were finally incubated in AM3, AM2-isocitrate, and AM2 liquid media. Only the third experiment led to growth in AM2-isocitrate after four days. Cultures grew overnight in AM3 (positive control) but no growth was observed after four days in AM2 (negative control). The four-day AM2-isocitrate culture (0.25 mL) was used to inoculate 10 mL of fresh AM2-isocitrate to enrich for recombinants able to grow on isocitrate. This second culture was streaked onto LB-glucose-Km plates to isolate putative recombinants. 4.4.3 Confirmation of the Δfrd::icd knockout strain. Double recombination of the Δfrd::icd construct in the A. succinogenes chromosome was confirmed by colony PCR. Of six putative recombinant colonies screened by PCR with primers P9 and P10, one yielded the expected 3.8-kb fragment (Figure 4.1). This putative mutant was confirmed using colony PCR screens with other primers. PCR with primers P11 and P12 confirmed that this colony did not contain any pCR2.1 sequences, while PCRs with primer pairs P23-P24 (internal to frdB) and P25-P26 (internal to E. coli icd) confirmed that this colony no longer contained the frdB gene but that it contained the icd gene (not shown). This mutant strain was called Δfrd::icd. Construction of the Δfrd::icd mutant was confirmed with in vitro ICD and fumarate reductase activity assays. Strain 130Z(pLGZ924) was the positive control for ICD activity. A. succinogenes 130Z showed almost no detectable ICD activity (0.30 ± 0.49 mU mg-1). With an 155 activity level of 28 ± 7 mU mg-1 Δfrd::icd showed over 90-fold higher ICD activity than 130Z, and 13-fold lower activity than 130Z(pLGZ924) (360 ± 130 U/mg). Higher ICD activity in 130Z(pLGZ924) reflects the fact that this strain contains multiple copies of icd. Strain Δfrd::icd lacks the frdAB genes and should be devoid of fumarate reductase activity. Indeed, Δfrd::icd showed over 500-fold lower fumarate reductase activity (0.0078 ± 0.0011 U mg-1) than 130Z (3.97 ± 0.08 U mg-1). 4.4.4 Excision of the selection marker. Strain Δfrd::icd was transformed with pCV933 (expressing the S. cerevisiae FLP recombinase) and plated on LB-glucose-Amp. One transformant colony was re-isolated on LBglucose-Amp and excision of the icd marker was tested by colony PCR using primers P25-P26. Out of the twenty potential Δfrd colonies tested, one yielded no PCR product. Colony PCR of this strain using primers P9-P10 yielded a 2.1-kb product instead of the 3.8-kb product obtained with Δfrd::icd, confirming that the icd gene was no longer present in that strain (Figure 4.1). The new strain was called Δfrd(pCV933). 4.4.5 Curing pCV933 from A. succinogenes Δfrd(pCV933). AO is commonly used to inhibit plasmid replication (33). After AO treatment, colonies isolated on LB-glucose-Km were replica-plated onto LB-glucose-Km and LB-glucose-Amp plates to identify Amps colonies. Twenty, ten, and ten Amps colonies from the Δfrd(pCV933) cultures grown with 50 μg ml-1, 75 μg ml-1, and 200 μg ml-1 AO, respectively, were screened by colony PCR with primers P21-P22. A single colony originating from the 50 μg ml-1 AO culture was confirmed to have lost the plasmid (not shown), and was called strain Δfrd. 156 A pCR2.1::ΔpflB::icd pflBup ppckA FRT A pCR2.1::ΔpflB::icd-600 FRT icd FRT ppckA FRT P11 A icd A P29 pCR2.1::ΔpflB::icd-400 ppckA 2.9 kb icd FRT A FRT P12 P32 P31 A pCR2.1::ΔpflB::icd-200 P11 P12 P30 A P11 pCR2.1 pflBdown FRT ppckA 2.5 kb icd P33 A FRT P34 P12 2.1 kb Figure 4.2 Physical maps of pCR2.1-ΔpflB::icd and its truncated constructs pCR2.1::ΔpflB::icd-600, pCR2.1::ΔpflB::icd-400, and pCR2.1::ΔpflB::icd-200. A: AscI; black triangles: primers; vertical arrows: USS sequences. 4.4.6 Construction of strain ΔpflB. The ΔpflB::icd construct contained a single USS in the frdBdown fragment, 33 nt downstream of the FRT-ppckA-icd-FRT cassette (Figure 4.2, top construct). XmnI-linearized pCR2.1-ΔpflB::icd, containing 2 kb of pCR2.1 DNA on each side of the knockout construct, was used to construct strain ΔpflB::icd.. Construction of the A. succinogenes ΔpflB::icd mutant was confirmed by colony PCR with primers P27-P28, which match sequences flanking the pflB deletion. Thirty-five of the thirty-six colonies tested yielded the expected 2-kb PCR fragment (Figure 4.3). PCR with primers P25-P26 confirmed that the putative mutants contained the E. coli icd gene (Figure 4.3). One isolated mutant was called ΔpflB::icd. The icd marker was excised by transforming strain ΔpflB::icd with pCV933. Of the thirty-eight colonies tested by PCR with primers P25-P26, eighteen yielded no PCR product. Excision of the icd gene was confirmed by colony PCR using primers P27-P28 (Figure 4.3), and one confirmed isolate was called ΔpflB(pCV933). 157 M M 11 2 2 3 3 44 3.0 kb 3.0 3.0 2.0 2.0 kb 2.0 1.5 1.5 kb 1.6 1.0 kb 1.0 1.0 A. succinogenes ΔpflB A. succinogenes ΔpflB::icd A. succinogenes 130Z pCR2.1-ΔpflB::icd A. succinogenes ΔpflB A. succinogenes ΔpflB::icd A. succinogenes 130Z pCR2.1-ΔpflB::icd Molecular weight markers P25-P26 Primers P25-P26 (A) (A) (kb) (kb) Gel images showing positively screened colonies in E.coli : Gel images showing positively screened colonies in E.coli : ΔpflB::icd-200 ΔpflB::icd-400 ΔpflB::icd-600 ΔpflB::icd-200 ΔpflB::icd-400 ΔpflB::icd-600 LL 1 1 2 32 4 3 L 1 2 L 1 2 L 1 L2 31 42 P27-P28 Primers P27-P28 (B) M 1 11 22 33 44 (kb) M 2 M 3 3 .0 2 .0 0.5 kb 0.5 0.5 L – 1 Kb Ladder; 3 – Negative Figure 4.3 Construction of strain ∆pflB (A) and natural transformation with the truncated L – 1 Kb Ladder; 3 – Negative products ΔpflB::icd-600, ΔpflB::icd-400, and ΔpflB::icd-200 (B). (A) Gel electrophoresis of the PCR products amplified from the ∆pflB::icd cassette (lane 1), from A. succinogenes 130Z (lane 2), from an A. succinogenes ∆pflB::icd derivative (lane 3), and from a ∆pflB(pCV933) derivative (lane 4) using primers P25-P26 (internal to icd) and P27-P28 (framing the pflB deletion). Lane M: DNA molecular markers. (B) Examples of colony PCRs after transformation with the truncated constructs ΔpflB::icd-600 (lane 1, PCR with primers P29-P30), ΔpflB::icd-400 (lane 2, PCR with primers P31-P32), and ΔpflB::icd-200 (PCR with primers P33-P34). Lane M: DNA molecular markers. Strain ΔpflB(pCV933) was cured of pCV933 by AO treatment. Ten and fifty μg mL-1 AO produced four and ten Amps colonies, respectively. The 100 μg mL-1 and 200 μg mL-1 AO treatments produced no Amps colonies. Of the fourteen Amps colonies, one from the 10 μg mL-1 AO and two from the 50 μg mL-1 AO treatments were shown to have lost the plasmid after two replicate colony PCR screens with primers P21 and P22 (not shown). One cured derivative from the 50 μg mL-1 AO treatment was called strain ΔpflB. During growth in AM3-25 mM NaHCO3, conditions that normally favor formate production (6, 29), ΔpflB did not produce any formate (Figure 4.4), confirming that this strain is devoid of pyruvate formate lyase activity. 158 unitsunits Absorption Absorption 0.08 0.07 0.08 0.06 0.07 0.05 0.06 0.04 0.05 0.03 0.04 0.02 0.03 0.01 0.02 0.00 0.01 Absorption units 4 2 4 Absorption units Absorption units B B 2 26 28 26 28 B 0.7 0.07 Pyr 0.08 0.6 B 0.06 0.07 0.5 0.05 0.06 0.4 0.04 0.05 0.3 0.03 0.04 0.2 0.02 0.03 0.01 0.1 0.02 0.0 0.00 0.01 8 10 2 412 6 8 10 12 0.00 Absorption units 0.00 A 0.07 0.8 0.08 130Z 130Z 130Z 0.06 A 0.7 0.07 0.05 0.6 0.06 0.04 0.5 0.05 0.03 Pyr Suc For Pyr Suc For 0.04 0.4 0.02 0.03 0.3 SucAce Pyr Suc For Ace Pyr For 0.01 0.02 0.2 Ace Ace 0.00 0.01 0.1 6 8 10 12 18 12 20 1422 1624 1826 2028 22 24 26 28 2 4 614 816 10 0.00 0.0 Minutes Minutes 10 124 14 146 16 16 8 18 20 12 22 14 2 12 16 28 18 20 22 24 6 88 10 18 1020 22 24 26 Time (min) Minutes Minutes Pyr Pyr ∆pflB 0.8 0.08 Absorption units 0.07 0.08 0.06 0.07 0.05 0.06 0.04 0.05 0.03 0.04 0.02 0.03 0.01 0.02 0.00 0.01 0.08 A A Absorption units unitsunits Absorption Absorption 0.08 Pyr Suc Suc Suc Suc Ace Ace Fum Ace Ace Fum Fum 14 8 16 10 18 20 14 22 16 6 12 18 2028 22 24 26 28 2 4 14 16 18 20 22 24 26 0.00 Time (min) Minutes Minutes 2 124 146 16 8 18 1020 12 16 28 18 20 22 24 2 4 6 8 10 22 14 24 26 Minutes Figure 4.4 UV spectra of the HPLC profiles of fermentation supernatants of A. Minutes Fum succinogenes 130Z, and ΔpflB grown on AM3 in the presence of 25 mM NaHCO3. Supernatant samples were collected immediately after inoculation (baseline) and after 12 h growth. Ace: acetate; For: formate; Fum: fumarate; Pyr: pyruvate. 4.4.7 Effects of the DNA construct length and of the length of the homologous regions on the efficiency of homologous recombination. The DNA fragments used in natural transformation so far contained approx. 1 kb of A. succinogenes DNA on each side of the selection marker for efficient recombination plus at least 1 kb of non-homologous plasmid DNA on each side to protect the homologous DNA from degradation by exonucleases. To test whether the non-homologous plasmid DNA on each side of 159 the recombination cassette could be omittted, the 3.8-kb Δfrd::icd cassette (amplified from pCR2.1-Δfrd::icd with primers P9-P10, Figure 4.1) was used in natural transformation. In this experiment, the selection marker was flanked by approx. 1 kb of A. succinogenes DNA on each side. Transformants grew in AM2-isocitrate. Forty percent of recombinant colonies screened by PCR with primers P9 and P10 yielded the expected 3.8-kb fragment (Table 4.2). This result indicates that no more than a total of 1 kb of DNA is needed on each side of the marker for recombination to take place. To determine what minimum length of homologous DNA is needed on each side of the selection marker to allow double recombination to take place, homologous recombination cassettes were generated that contained 200 bp, 400 bp, and 600 bp homologous DNA on each side of the selection marker. pCR2.1-ΔpflB::icd was used as the template because truncating the frdup and frdCD fragments would have deleted both USS sequences (Figure 4.1). The ΔpflB::icd600, -400 and -200 fragments were amplified from pCR2.1-ΔpflB::icd with primers P29-P30, P31-P32, and P33-P34, respectively, and cloned back into pCR2.1 (Figure 4.2). The fragments used for transformation were generated by PCR with primers P11-P12, including 1 kb of vector DNA on each side of the recombination cassette. Three independent natural transformations were performed with each fragment, in which 15 colonies were screened by PCR. While the standard deviations were large in all cases, averages of 47%, 57%, and 60% colonies contained the ΔpflB::icd deletion for the transformations with fragments ΔpflB::icd-600, -400, and -200, respectively. These results suggest that 200 bp of homologous DNA on each side of the selection marker are still enough to allow double recombination. Conversely, all transformations using the ΔpflB::icd-600 cassette amplified with primers P30-P31, the ΔpflB::icd-400 amplified with 160 primers P31-P32, or the ΔpflB::icd-200 cassette amplified with primers P33-P34 yielded no growth in AM2-isocitrate and no ΔpflB::icd deletions. 4.4.8 Construction of double knockout mutants. To determine whether the same knockout strategy could be used to introduce two successive deletions in the same strain, three genes deemed non-essential for growth in AM3 glucose were targeted for deletion in strain ΔpflB. The genes encoding citrate lyase and aconitase were chosen as targets, because the tricarboxylic acid cycle is incomplete in A. succinogenes. The third target gene was lacZ. The strategy used to build the three deletions was identical to that used to build the Δfrd and ΔpflB strains. Citrate lyase is encoded by three genes organized in an operon, citDEF (Asuc_1194-1196). CitE encodes the subunit with lyase activity. The citrate lyase deletion, Δcit, encompassed over half of citD, the entire citE, and over a third of citF (not shown). One USS was present in the upstream fragment. The 2.6-kb aconitase gene (acn, Asuc_0185) is immediately preceded by the only USS in that area. To preserve that USS and to preserve the promoter of the gene downstream of acn, the acn deletion encompassed only 1.6 kb inside the acn gene. The lacZ deletion encompassed the entire 3.0-kb lacZ gene, including the only four USSs present in the area. For this reason, a USS was introduced into the reverse nested primer used to fuse the lacZ up and down regions (P40, Table S4.1). The three double knockout mutant strains were obtained with frequencies ranging between 25% and 100% (Table 4.2). Note that double recombinants were obtained with frequencies of 25% or higher for Δcit::icd in strains ΔpflB and 130Z and for Δacn::icd in ΔpflB with only 0.52 kb to 0.6 kb of total flanking DNA on each side of the selection marker (Table 4.2). 161 Table 4.2 Frequencies of knockout mutations introduced by natural transformation in A. succinogenes strains 130Z and ΔpflB Homologous DNA a Vector DNA on each Knockout frequency Transformation Host on each side of the side of the (positive cassette strain selection cassette recombination colonies/colonies (kb) cassette (kb) screened) Knockout frequency (%) Δfrd::icd 130Z 1 0.8 1/6 17 Δfrd::icd 130Z 1 0 6/15 40 Δfrd::icd 130Z 1 0.8 No growtha – ΔpflB::icd 130Z 1 1.9 35/36 97 ΔpflB::icd-600 130Z 0.6 1.0 6/16, 15/15, 6/15 60 ± 35 ΔpflB::icd-400 130Z 0.4 1.0 4/15, 7/15, 15/15 57 ± 38 ΔpflB::icd-200 130Z 0.2 1.0 3/15, 4/15, 14/15 47 ± 41 ΔpflB::icd-600 130Z 0.6 0 No growth – ΔpflB::icd-400 130Z 0.4 0 No growth – ΔpflB::icd-200 130Z 0.2 0 No growth – Δcit::icd 130Z 0.5 1.0 5/12 42 Δcit::icd 130Z 0.5 0.1 3/12 25 ΔlacZ::icd ΔpflB 0.8 1.0 17/19 89 Δcit::icd ΔpflB 0.5 1.0 30/30 100 Δcit::icd ΔpflB 0.5 0.1 6/12 50 Δacn::icd ΔpflB 0.42 0.1 4/16 25 Experiment repeated five times with the same results. Figure 4.5 confirms that ΔpflB ΔlacZ::icd transformants no longer have β-galactosidase activity. ΔpflB Δcit showed 27-fold less in vitro citrate lyase activity than ΔpflB (0.02 μmol min-1 mg-1 in ΔpflB Δcit vs. 0.55 μmol min-1 mg-1 in ΔpflB), confirming the double deletion. 162 ΔpflB ΔpflB-ΔlacZ::icd Figure 4.5 Phenotypic characterization of double knockout mutant ∆pflB-∆lacZ::icd. One colony each of ∆pflB and one ∆pflB-∆lacZ::icd were patched on LB X-gal Amp plates. 4.4.9 Development of a selective solid medium. A. succinogenes 130Z grows slowly on AM3 plates, with colonies visible only after 4 days. Even though it contains multiple copies of the E. coli icd gene, 130Z(pLGZ924) did not form visible colonies on AM2-isocitrate plates. To develop a selective agar-isocitrate medium that allows strains containing a single copy of E. coli icd to form colonies, all amino acids but glutamate, glutamine, arginine, and proline (i.e., amino acids derived from glutamate) were added to the medium. Glutamate was replaced by 30 mM DL-isocitrate. 130Z(pLGZ924) formed colonies on this medium after two days, but Δfrd::icd did not, even after six days. The same medium containing additional vitamins, cysteine, and methionine (i.e., medium AM16-isocitrate) supported growth of Δfrd::icd, with colonies forming after five days. AM16-isocitrate plates were tested as selective solid medium in two replicate natural transformations of strain 130Z with 1 μg NcoI-linearized pCR2.1-Δfrd::icd. Transformation mixtures were inoculated into liquid AM3 and AM2-isocitrate (250 µL in each), and spread on AM16-isocitrate plates (225 µL spread at 100, 10-1, and 10-2 dilutions). As expected, transformants grew overnight in AM3 and after four days in liquid AM2-isocitrate. Colonies grew on AM16-isocitrate plates after five days, with an efficiency of approximately 140 colonies/μg DNA. Eighteen colonies grown on AM16-isocitrate plates were tested by colony PCR using primers P9-P2. All were Δfrd::icd 163 mutants, confirming that AM16-isocitrate plates can be used to directly select icd-containing knockout mutant constructs in A. succinogenes. All attempts to measure transformation and double recombination efficiencies failed because of a yet unexplained problem with serial dilutions and plating. Starting from fresh dense precultures, all A. succinogenes strains tested, including 130Z, grew as lawns when plated from the undiluted preculture, they grew as numerous compact colonies when plated from the 10-1 dilution, but they produced only a few colonies when plated from the 10-2 dilution and no colonies at higher dilutions. Similar results were observed when the serial dilutions were prepared with AM3 phosphate buffer, AM3 with and without glucose, LB, or with the supernatant of the preculture, and when the bacterial suspensions were plated on LB-glucose, AM3, AM17, or AM16-isocitrate. 4.4.10 Knockout mutants via electroporation. The knockout mutants obtained using natural transformation showed that linear DNA fragments could recombine with the A. succinogenes chromosome. Despite previous failed attempts at constructing knockout mutants using electroporation and antibiotic resistance genes as selection markers, electroporation was tested again as an alternative transformation method to create knockout mutants in A. succinogenes. Electroporations of strain 130Z with NcoI-linearized pCR2.1-Δfrd::icd were performed using 0.1‒1 µg DNA in 100 ng increments. After recovery, electroporation mixtures were washed once with MIV and resuspended in MIV. One third of each electroporation mixture was inoculated in liquid AM3, one third was inoculated in liquid AM2-isocitrate, and one third was spread on a single AM16-isocitrate plate. Electroporations with 300 ng or more linearized pCR2.1-Δfrd::icd 164 consistently yielded colonies on AM16-isocitrate plates. All colonies growing on AM16isocitrate plates that were tested by colony PCR with primers P9-P2 were confirmed as Δfrd::icd knockouts, showing that electroporation can be used as an alternative to natural transformation to introduce linear DNA in A. succinogenes for homologous recombination. 4.5 Discussion The fermentative metabolism of A. succinogenes has been studied in detail using fermentation balances, in vitro enzyme activity assays, genome sequencing, and metabolic flux analysis, and how ATP is produced and NAD+ is regenerated in the different fermentative branches is well understood (6, 17, 21, 22, 29). To enhance the production of a given fermentative metabolite (e.g., fumarate or succinate), we needed a means to produce knockout mutants. Here, we demonstrated that natural transformation and electroporation can be used to introduce DNA in A. succinogenes for recombination, that the E. coli icd gene can be used as a positive selection marker for recombination events, that the yeast Flp/FRT recombination system can be used to excise the selection marker, that plasmids can be cured from A. succinogenes using AO, and that the selection marker can be reused to introduce at least two consecutive deletions into the same strain. The 64-nt scar (FRT sequence flanked by AscI sites) interrupting the deletion does not by itself leave the remaining sequences in frame, but if an in-frame deletion is desired, the deletion itself can be designed with a frameshift to place the remaining sequences back in frame. Except for our first transformation to construct the Δfrd::icd strain, at least 25% of colonies obtained after all other natural transformations were knockout mutants (Table 4.2). The frequency of knockout colonies increased to 100% when transformants (obtained by 165 electroporation) were selected directly on the enriched defined medium, AM16-isocitrate. Note that we could not repeat the natural transformation with the Δfrd::icd cassette in at least five more attempts, suggesting that the efficiency of the method is somewhat gene-specific. The sequences of the two USS repeats present in the Δfrd::icd cassette were closer to the consensus USS for A. succinogenes than the single USS in the ΔpflB::icd cassette was (not shown). Therefore poor recognition of the USSs flanking the Δfrd::icd cassette by the DNA uptake machinery cannot explain these results. We have also tried repeatedly to knockout other genes (i.e., ackA, pta, pykA, and zwf) in 130Z and ΔpflB without success, but these genes might be essential for growth, or we have not yet found conditions in which a knockout strain can grow. With the ΔpflB::icd, ΔlacZ::icd, Δcit::icd, and Δacn::icd constructs, we demonstrated that a single USS in the knockout construct is enough to allow DNA uptake by A. succinogenes. Introducing a USS in the ΔlacZ::icd construct―which would otherwise not have contained any USS, also allowed double recombination to take place. Linear constructs of varying lengths could be used for transformation, from the entire plasmid carrying the knockout construct (e.g., ΔpflB::icd in 130Z) down to recombination cassettes containing 0.52 kb to 0.6 kb of DNA on each side of the selection marker (e.g., Δacn::icd in ΔpflB) (Table 4.2). While flanking regions of 0.5 kb to 0.6 kb on each side of the cassette yielded knockouts with frequencies as low as 25% (or even 0% for ΔpflB::icd-600), 1 kb of flanking DNA on each side was enough to protect against exonucleases prior to recombination, routinely yielding knockout frequencies between 42% and 100% (Table 4.2). As little as 200 bp of homologous DNA on each side of the selection marker was enough for efficient double recombination, as long as the homologous DNA was itself flanked by non-homologous DNA on each side. 166 The development of a markerless knockout method for A. succinogenes will greatly facilitate future genetic engineering of A. succinogenes. Strain ΔpflB represents an important first step in engineering a strain that can produce succinate at near maximum theoretical yields. 4.6 Acknowledgements This work was supported by a grant from the Michigan Economic Development Corporation, by MSU startup funds, and by grant # 2010-04061 from the USDA National Institute for Food and Agriculture’s Sustainable Bioenergy Research Program to CV. BS was supported in part by a research fellowship from the MSU Quantitative Biology Initiative. NM was supported in part by a recruiting fellowship from the MSU Plant Research Laboratory. KT was supported by a Khorana scholarship. We thank Reena Jain, Jean Kim, and Maeva Bottex for their technical assistance, as well as Dr. Michael Bagdasarian for his critical reading of the manuscript. 167 APPENDIX 168 Table A4.1 Oligonucleotide primers used in this study Primer P1 P2 P3 P4 P5 P6 P7 P8 P9 P10 P11 P12 P13 P14 P15 P16 P17 P18 P19 P20 P21 P22 P23 P24 P25 P26 P27 P28 P29 P30 P31 P32 P33 P34 P35 P36 P37 P38 P39 Sequencea (restriction site) GTTCTAGAGATGGAAAGTAAAGTAGTTG (XbaI) CTGAATTCATTACATGTTTTCGATGATC (EcoRI) GAGGCGCGCCGAAGTTCCTATTCCGAAGTTCCTATTCTCTAGAA AGTATAGGAACTTCTCGATAAATTGAAAATGCAGCAA (AscI) CTGGCGCGCCGAAGTTCCTATACTTTCTAGAGAATAGGAACTTC GGAATAGGAACTTCTTACATGTTTTCGATGATCGC (AscI) TTAGGTACCGGCAACAAAGG CCCCTATGCTTGGCGCGCCCGTAAACCTAAACGCGCAAT (AscI) GGCGCGCCAAGCATAGGGGGAAAGCAAT (AscI) GTCCGATTTGGGTTTTGCTA GTAAGCTTACGGCAAACACGATCACATA CACTCGAGGCACCGCCTGTCACTAAAAT CCGGATCAAGAGCTACCAAC CGAAACGATCCTCATCCTGT CGTTAACCGTGGGAATCAGT TTACGTTACCCCAGGCGCGCCCTTCCTTTTGCTAGTATTGATAAT GA (AscI) GGCGCGCCTGGGGTAACGTAATAAAAATGTAATG (AscI) TCTCTCCTTCGCGGAATAAA TGAGCCTGACTGGTAAATCCA CACATCGACCCCGATAACTT ATGAGGTGATCTAGATGCCACAATTTGGTATATTATGTAAA (XbaI) CGGCCAGTGAATTCGAGCTCTTATATGCGTCTATTTATGTAGG (SacI) CGTTGTAAAACGACGGCC AATTTTAAATTATCAATGAGGTG TGCGTTACAACCCTGAAACA TCTTTCGCACTTTCCAGCTT GAAGGTGATGGAATCGGTGT GTTACGGTTTTCGCGTTGAT AATTTGCGGATCTGGATGTC TGGTGTATTCCGTCAGTTCG GTTGCGAACTGTTCACCTC CAATATCGCGGCGTTCCAGC GCATTGGGCATTTTGTGTAAC CTAGTAGAAATGCGCGTTTTA GAGTTTTGGCAGGCAATC CTTTCCCAGCGTTCCACAAAAAC GGATAAACTCGGCGTACG GCGGTTATGGGCGCGCCGCAGAATCATGCTGAACTCC (AscI) GGCGCGCCCATAACCGCATAAAAATACAGGGCAG (AscI) CATGAAACCGAACAGTATTGTGC CTTGCCAAACCGACCGAAAG 169 Specificityb (directionc) E. coli K-12 icd gene (F) E. coli K-12 icd gene (R) AscI-FRT- ppckA (F) E. coli icd-FRT-AscI (R) frdup (F) frdup (R) frdCD (F) frdCD (R) Δfrd-fusion, nested (F) Δfrd-fusion, nested (R) pCR2.1 backbone (F) pCR2.1 backbone (R) pflBup (F) pflBup (R) pflBdown (F) pflBdown (R) ΔpflB-fusion, nested (F) ΔpflB-fusion, nested (R) S. cerevisiae flp (F) S. cerevisiae flp (R) pLGZ920 (F) pLGZ920 (R) frdB (F) frdB (R) Inside E. coli K-12 icd (F) Inside E. coli K-12 icd (R) 217 nt upstream of pflB (F) 146 nt downstream of pflB (R) ΔpflB::icd-600 (F) ΔpflB::icd-600 (R) ΔpflB::icd-400 (F) ΔpflB::icd-400 (R) ΔpflB::icd-200 (F) ΔpflB::icd-200 (R) lacZup (F) lacZup (R) lacZdown (F) lacZdown (R) ΔlacZ-fusion, nested (F) Table A4.1 (cont’d) TATTGATAATGAAAATCCGACCGCACTTGGCAGTACCGGCGTAT ΔlacZ-fusion, nested (R) TCCTCd P41 CTTGACGGTAGCAGTGATC citDEFup (F) P42 GGCGCGCCGAATTGCTTTCCGAC (AscI) citDEFup (R) P43 GAAAGCAATTCGGCGCGCCAATCAGTATCGGCCAGG (AscI) citDEFdown (F) P44 CCGTATCGCAGTGACCGC citDEFdown (R) P45 CTGGTGGTGACGATCATAC ΔcitDEF-fusion, nested (F) P46 GGTATCGATTTCCAATGCGG ΔcitDEF-fusion, nested (R) P47 CGGCGAAGTGTTCTACGATG acnup (F) P48 GGCGCGCCGTGGTATCCACCTCAAC (AscI) acnup (R) P49 GGATACCACGGCGCGCCATATGCCACAGCACGG (AscI) acndown (F) P50 CATGCGGTAGAATTACTCGG acndown (R) P51 ACGATTATCCGAGTCATGAG Δacn-fusion, nested (F) P52 GCCAGATTCGCTCCGCTTG Δacn-fusion, nested (R) P53 TGTAAAACGACGGCCAGT M13 forward P54 CAGGAAACAGCTATGACC M13 reverse a 5’ to 3’ direction; restriction sites are underlined and named at the end of the primer sequences b If unspecified, gene names refer to A. succinogenes c F, forward primer; R, reverse primer d Primer P41 contains the USS ACCGCACTT P40 170 kb M 1 2 3 4 M 5 6 M 7 8 3.0 2.0 1.6 1.0 0.5 Figure A4.1 Verification of the construction of strains ∆pflB ∆cit and ∆pflB ∆acn by PCR. Lanes M: DNA molecular markers. PCR products amplified from 130Z genomic DNA (Lane 1), ∆pflB genomic DNA (Lanes 2, 5, and 7), a ∆pflB ∆acn colony (Lanes 3 and 6), and a ∆pflB ∆cit colony (Lanes 4 and 8). Lanes 1 to 4: verification that the host for the double mutations is strain pflB; primers P27-P28 (flanking pflB) amplify a 2.7-kb fragment in 130Z, and a 434-bp fragment in ∆pflB, ∆pflB ∆acn, and ∆pflB ∆cit). Lanes 5 and 6: verification that acn is deleted in strain ∆pflB ∆acn; primers P51-P52 amplify a 2.4-kb fragment in ∆pflB and a 855-bp fragment in ∆pflB ∆acn. Lanes 7 and 8: verification that citDEF are deleted in strain ∆pflB ∆cit; primers P55P56 amplify a 2.2-kb fragment in ∆pflB and a 524-bp fragment in ∆pflB ∆cit. The PCR products in lanes 6 and 8 would be 2.35 kb (∆acn) and 2.0 kb (∆cit), if the deletions still contained the icd marker. 171 REFERENCES 172 REFERENCES 1. Werpy TA, Petersen G. 2004. Top value added chemicals from biomass for the U.S. Department of Energy (DOE) by National Renewable Energy Laboratory (NREL). 2. Bozell JJ, Petersen G. 2010. Technology development for the production of biobased products from biorefinery carbohydrates—the US Department of Energy’s Top 10 revisited. Green Chemistry 12:539–554. 3. Guettler MV, Jain MK, Rumler D. 1996. Method for making succinic acid, bacterial variants for use in the process, and methods for obtaining variants. U.S. patent 5,573,931. 4. Guettler MV, Jain MK, Soni BK. 1996. Process for making succinic acid, microorganisms for use in the process and methods of obtaining the microorganisms. U.S. patent 5,504,004. 5. Guettler MV, Rumler D, Jain MK. 1999. Actinobacillus succinogenes sp. nov., a novel succinic-acid-producing strain from the bovine rumen. Int J Syst Bacteriol 49:207–216. 6. van der Werf MJ, Guettler MV, Jain MK, Zeikus JG. 1997. Environmental and physiological factors affecting the succinate product ratio during carbohydrate fermentation by Actinobacillus sp. 130Z. Arch Microbiol 167:332–342. 7. Kurzrock T, Weuster-Botz D. 2010. Recovery of succinic acid from fermentation broth. Biotechnol Lett 32:331–339. 8. Kim P, Laivenieks M, McKinlay J, Vieille C, Zeikus JG. 2004. Construction of a shuttle vector for the overexpression of recombinant proteins in Actinobacillus succinogenes. Plasmid 51:108–115. 9. Oswald W, Tonpitak W, Ohrt G, Gerlach G-F. 1999. A single-step transconjugation system for the introduction of unmarked deletions into Actinobacillus pleuropneumoniae serotype 7 using a sucrose sensitivity marker. FEMS Microbiol Lett 179:153–160. 10. Tatum FM, Briggs RE. 2005. Construction of in-frame aroA deletion mutants of Mannheimia haemolytica, Pasteurella multocida, and Haemophilus somnus by using a new temperature-sensitive plasmid. Appl Environ Microbiol 71:7196–7202. 11. Kim JM, Lee KH, Lee SY. 2008. Development of a markerless gene knock-out system for Mannheimia succiniciproducens using a temperature-sensitive plasmid. FEMS Microbiol Lett 278:78–85. 12. Maughan H, Sinha S, Wilson L, Redfield R. 2008. Competence, DNA uptake, and transformation in Pasteurellaceae, p 79–98. In Kuhnert P, Christensen H (ed), Pasteurellaceae Biology, Genomics and Molecular Aspects. Caister Academic Press. 173 13. Redfield RJ, Findlay WA, Bossé J, Kroll JS, Cameron AD, Nash JH. 2006. Evolution of competence and DNA uptake specificity in the Pasteurellaceae. BMC Evol Biol 6:82. 14. Sisco KL, Smith HO. 1979. Sequence-specific DNA uptake in Haemophilus transformation. Proc Natl Acad Sci USA 76:972–976. 15. Danner DB, Deich RA, Sisco KL, Smith HO. 1980. An 11-base pair sequence determines the specificity of DNA uptake in Haemophilus transformation. Gene 11(3– 4):311–318. 16. Wang Y, Orvis J, Dyer D, Chen C. 2006. Genomic distribution and functions of uptake signal sequences in Actinobacillus actinomycetemcomitans. Microbiology 152:3319– 3325. 17. McKinlay JB, Laivenieks M, Schindler BD, McKinlay AA, Siddaramappa S, Challacombe JF, Lowry SR, Clum A, Lapidus AL, Burkhart KB, Harkins V, Vieille C. 2010. A genomic perspective on the potential of Actinobacillus succinogenes for industrial succinate production. BMC Genomics 11:680. 18. Redfield RJ, Cameron AD, Qian Q, Hinds J, Ali TR, Kroll JS, Langford PR. 2005. A novel CRP-dependent regulon controls expression of competence genes in Haemophilus influenzae. J Mol Biol 347:735–747. 19. Cox MM. 1983. The FLP protein of the yeast 2-microns plasmid: expression of a eukaryotic genetic recombination system in Escherichia coli. Proc Natl Acad Sci USA 80:4223–4227. 20. Zhu XD, Pan G, Luetke K, Sadowski PD. 1995. Cleavage-dependent ligation by the FLP recombinase. Characterization of a mutant FLP protein with an alteration in a catalytic amino acid. J Biol Biol 270:11646–11653. 21. McKinlay JB, Shachar-Hill Y, Zeikus JG, Vieille C. 2007. Determining Actinobacillus succinogenes metabolic pathways and fluxes by NMR and GC-MS analyses of 13Clabeled metabolic product isotopomers. Metab Engin 9:177–192. 22. McKinlay JB, Vieille C. 2008. 13C-metabolic flux analysis of Actinobacillus succinogenes fermentative metabolism at different NaHCO3 and H2 concentrations. Metab Engin 10:55–68. 23. Sawers RG, Clark DP. 2004. Chapter 3.5.3, Fermentative pyruvate and acetylcoenzyme A metabolism. In Böck A, Curtiss III R, Kaper JB, Karp PD, Neidhardt FC, Nyström T, Slauch JM, Squires CL, Ussery D (ed), EcoSal—Escherichia coli and Salmonella: Cellular and Molecular Biology. ASM Press, Washington, DC. 174 24. Stols L, Donnelly MI. 1997. Production of succinic acid through overexpression of NAD+-dependent malic enzyme in an Escherichia coli mutant. Appl Environ Microbiol 63:2695–2701. 25. Jantama K, Haupt MJ, Svoronos SA, Zhang XL, Moore JC, Shanmugam KT, Ingram LO. 2008. Combining metabolic engineering and metabolic evolution to develop nonrecombinant strains of Escherichia coli C that produce succinate and malate. Biotechnol Bioengin 99:1140–1153. 26. Zhang X, Jantama K, Shanmugam KT, Ingram LO. 2009. Reengineering Escherichia coli for succinate production in mineral salts medium. Appl Environ Microbiol 75:7807– 7813. 27. Lee SJ, Song H, Lee SY. 2006. Genome-based metabolic engineering of Mannheimia succiniciproducens for succinic acid production. Appl Environ Microbiol 72:1939–1948. 28. Ausubel FM, Brent R, Kingston RE, Moore DD, Seidman JG, Smith JA, Struhl K (ed). 1993. Current Protocols in Molecular Biology. Greene Publishing & WileyInterscience, New York. 29. McKinlay JB, Zeikus JG, Vieille C. 2005. Insights into the Actinobacillus succinogenes fermentative metabolism in a chemically defined growth medium. Appl Environ Microbiol 71:6651–6656. 30. Choi KH, Kumar A, Schweizer HP. 2006. A 10-min method for preparation of highly electrocompetent Pseudomonas aeruginosa cells: application for DNA fragment transfer between chromosomes and plasmid transformation. J Microbiol Methods 64:391–397. 31. Poje G, Redfield RJ. 2003. Transformation of Haemophilus influenzae. Methods Mol Med 71:57–70. 32. Bosse JT, Sunita S, Schippers T, Kroll JS, Redfield RJ, Langford PR. 2009. Natural competence in strains of Actinobacillus pleuropneumoniae. FEMS Microbiol Lett 298:124–130. 33. Trevors JT. 1986. Plasmid curing in bacteria. FEMS Microbiol Rev 32:149–157. 175 Chapter 5 Respiratory glycerol metabolism of Actinobacillus succinogenes 130Z for succinate production 176 The work completed in this chapter is part of a collaboration with Dr. Bryan Schindler (primary author, a former member of the Vieille lab) and Dr. Claire Vieille, published as: Schindler BD, Joshi RJ, Vieille C. 2014. Respiratory glycerol metabolism of Actinobacillus succinogenes 130Z for succinate production. J Ind Microbiol Biotechnol 41:1339-1352. doi: 10.1007/s10295-014-1480-x Article can be viewed at the following URL: http://rdcu.be/v6Wy My contribution to this report included:  Developing and optimizing a microaerobic continuous culture system for A. succinogenes (chemostats) for growth on glycerol.  Conducting continuous culture experiments with A. succinogenes 130Z and carrying out fermentation balances.  Drafting portions of the manuscript for above bullet points Reading and reviewing entire manuscript. 177 Chapter 6 Conclusions and future directions 178 6.1 Introduction A. succinogenes is one of the best natural succinate producers known. Many advances have been carried out with respect to its metabolism on substrates like glucose, glycerol, and lignocellulosic hydrolysates (1-4). Microorganisms such as E. coli, S. cerevisiae, and C. glutamicum are being studied and engineered for succinate production as well. Many of the succinate production increases obtained in E. coli were obtained by mimicking A. succinogenes metabolism. Despite much insight into A. succinogenes’s metabolism, we still know very little about other aspects, such as succinate efflux transporters or small RNAs (sRNAs) in A. succinogenes. In this dissertation, we have focused on identifying succinate exporters, and overexpressing four individual transporters increased succinate production. Additionally, we have also identified sRNAs in A. succinogenes and shown that synthetic sRNAs can be used in A. succinogenes for inhibiting expression of unwanted pathways. The knowledge gained about succinate transporters and sRNAs in this dissertation can be used for future metabolic engineering of A. succinogenes for succinate production. 6.2 Succinate transporters in A. succinogenes We have identified succinate transporters in A. succinogenes using a combination of proteomics and RNA sequencing (RNAseq). The transporters with the most hits in our proteomics studies and high transcripts levels were different from the succinate exporters identified in E. coli and C. glutamicum. Although homologs of the E. coli and C. glutamicum succinate exporters exist in A. succinogenes, none of them were among the top hits in our proteomics and transcriptomics studies. Asuc_1999, Asuc_0142, Asuc_2058, and Asuc_1990-91 were the transporter candidates that topped our proteomics list. Since expressing the transporters 179 under the control of a strong promoter was toxic for A. succinogenes, we developed a promoter library with a large span of expression strength. Expressing Asuc_1999 under the control of weaker promoter ppckA-103 increased the succinate yield in A. succinogenes by 24%, mainly at the cost of biomass production. 6.3 Identification of sRNAs in A. succinogenes The purpose of this study was to gain some insight into the size, structure, and function of sRNAs in A. succinogenes, especially of Hfq-dependent sRNAs. We identified 260 sRNAs in A. succinogenes in glucose- and glycerol-grown cultures, out of which 24 had predicted Rhoindependent terminators, a characteristic feature of Hfq-dependent sRNAs. We were also able to validate 14 randomly selected sRNAs by RT-PCR. One of the major goals of this study was to use the knowledge gained from the RNAseq studies of sRNAs and design a synthetic sRNA to inhibit the translation of select mRNAs in A. succinogenes. We were most interested in this approach since knockout mutants maybe impossible to construct for certain essential genes, and this approach offers a much more tunable way to lower gene expression. In this study, we successfully designed synthetic sRNAs targeting β-galactosidase and acetate kinase expression. We were able to decrease β-galactosidase activity by 32% and to decrease the acetate yield by 14%. 180 6.4 Future directions 6.4.1 Experimental identification of Hfq-binding sRNAs In this study we used computational and manual approaches to identify possible Hfqbinding sRNAs. We still need to experimentally verify these Hfq-binding sRNAs, though. One approach to identifying Hfq-binding sRNAs as well as their target mRNAs would be to compare transcriptomics data between a Δhfq strain and the wild-type strain. I built a Δhfq::CmR strain, but that strain was only transiently culturable, not long enough to allow transcriptomic analysis. Another approach is copurification using a strain expressing His-tagged Hfq, which would allow us to isolate and sequence Hfq-binding sRNAs. This approach would allow us to confirm a number of the predicted Hfq-binding sRNAs in our RNAseq studies, as well as identify some mRNAs regulated by Hfq-dependent sRNA. Because we could not complement our Δhfq::CmR strain with a plasmid-borne His-tagged hfq gene, we are copurifying His6-tagged Hfq with its binding RNAs after expression in wild-type A. succinogenes instead. Results from this experiment will yield additional scaffold sequences from Hfq-binding sRNAs. 6.4.2 Continuing the development of synthetic sRNAs in A. succinogenes Once we confirm Hfq-binding sRNAs experimentally, it will be possible to compare many scaffolds sequences and build additional synthetic sRNAs. Additionally, we need more studies with synthetic sRNAs designed with differing scaffold and targeting sequence lengths to modulate the binding with and regulation of the target mRNA. A more in depth look at the secondary structures of the native Hfq-binding sRNAs is required. Mimicking the secondary structure of native Hfq-binding sRNAs in our synthetic sRNAs may increase stability and 181 improve effectiveness of the designed sRNAs. All these studies will help in building a library of synthetic sRNA scaffolds, where simply replacing the target binding regions would be possible. 6.4.3 Identifying the major succinate transporter(s) in A. succinogenes In our proteomics and RNAseq studies we were able to identify a few transporter candidates that may be responsible for exporting succinate out of the cell. However, we were unable to identify one major succinate transporter that functions under anaerobic conditions. As seen in E. coli and C. glutamicum, several succinate exporters may contribute to succinate efflux under anaerobic conditions. Our attempts to knockout any of the four succinate exporters identified in this study have been unsuccessful so far. But once we have these knockout mutants, we will be able to test the knockout strains under anaerobic and microaerobic conditions to shed more light on succinate exporters active under different conditions in A. succinogenes. 6.4.4 Construction of a ∆ackA knockout mutant All our attempts at making a ∆pflB∆ackA mutant have been unsuccessful so far. Recently, the Bechkam group was able to make a ∆pflB∆ackA mutant strain using homologous recombination (5). The difference in approach was the use of a different selection marker and growth medium. The authors used an antibiotic resistance gene as their selection marker and tryptic soy broth supplemented with glucose and the antibiotic as the selection medium. In the same study, the authors overexpressed phosphoenolpyruvate carboxykinase, malate dehydrogenase, and fumarase independently and in the same strain. The succinate yields and titers of strains expressing malate dehydrogenase and all the enzymes together were higher than 182 that of the wild-type strain. However, the same constructs in the ∆pflB∆ackA background did not cause any increase in succinate titer or yield. Our lack of success in constructing a ∆ackA strain suggests that the acetate pathway is essential for A. succinogenes’s growth on a minimal medium such as AM3. We have previously constructed ∆lacZ::CmR and Δhfq::CmR strains by selecting colonies on LB medium supplemented with chloramphenicol, so we should be able to obtain a ∆ackA mutant using the same approach. 6.4.5 Combination of different approaches to increase succinate production Evolved strains of A. succinogenes that produce more succinate than the wild-type strain on glucose and xylose have been isolated in our lab. The X strains, evolved for fast growth on xylose, have mutations in the xylE and fusA genes, encoding the xylose permease and elongation factor G, respectively. Introducing the xylE mutation in the wild-type strain increased the succinate yield by 40%. We have a ΔpflB mutant (devoid of pyruvate formate lyase) that no longer produces formate and produces 69% more succinate than the wild-type strain on glucose (6). We also have identified succinate exporters that, upon overexpression, increase succinate production as well. And we now have a library of promoters of a range of strengths, Combining different mutations, expressing succinate exporter(s) as well as possibly some of the enzymes in the succinate pathway should allow us to engineer strains that produce record amounts of succinate with minimum byproducts. 183 REFERENCES 184 REFERENCES 1. McKinlay JB, Shachar-Hill Y, Zeikus JG, Vieille C. 2007. Determining Actinobacillus succinogenes metabolic pathways and fluxes by NMR and GC-MS analyses of 13Clabeled metabolic product isotopomers. Metab Eng 9:177–192. 2. McKinlay JB, Vieille C. 2008. 13C-metabolic flux analysis of Actinobacillus succinogenes fermentative metabolism at different NaHCO3 and H2 concentrations. Metab Eng 10:55–68. 3. Schindler BD, Joshi RJ, Vieille C. 2014. Respiratory glycerol metabolism of Actinobacillus succinogenes 130Z for succinate production. J Ind Microbiol Biotechnol 41:1339-1352. 4. McPherson N. 2017. Novel Insights Into Sugar and Succinate Metabolism of Actinobacillus succinogenes from Strains Evolved for Improved Growth on Lignocellulose Hydrolysate Sugars. Doctoral. Michigan State University. 5. Guarnieri MT, Chou YC, Salvachua D, Mohagheghi A, St John PC, Peterson DJ, Bomble YJ, Beckham GT. 2017. Metabolic Engineering of Actinobacillus succinogenes Provides Insights into Succinic Acid Biosynthesis. Appl Environ Microbiol 83: e0099617. 6. Schindler B. 2011. Understanding and improving respiratory succinate production from glycerol by Actinobacillus succinogenes. Doctoral Dissertation. Michigan State University. 185