SUNY College of Environmental Science and Forestry Digital Commons @ ESF

Dissertations and Theses

Spring 4-15-2020

Enhancing polyhydroxyalkanoate biosynthesis in : A genetic engineering and process optimization approach towards functionalized polymeric nanomedicine

Ryan Scheel SUNY College of Environmental Science and Forestry, [email protected]

Follow this and additional works at: https://digitalcommons.esf.edu/etds

Part of the Environmental Engineering Commons

Recommended Citation Scheel, Ryan, "Enhancing polyhydroxyalkanoate biosynthesis in Escherichia coli: A genetic engineering and process optimization approach towards functionalized polymeric nanomedicine" (2020). Dissertations and Theses. 161. https://digitalcommons.esf.edu/etds/161

This Open Access Dissertation is brought to you for free and open access by Digital Commons @ ESF. It has been accepted for inclusion in Dissertations and Theses by an authorized administrator of Digital Commons @ ESF. For more information, please contact [email protected], [email protected]. ENHANCING POLYHYDROXYALKANOATE BIOSYNTHESIS IN ESCHERICHIA

COLI: A GENETIC ENGINEERING AND PROCESS OPTIMIZATION APPROACH

TOWARDS FUNCTIONALIZED POLYMERIC NANOMEDICINE

By

Ryan A. Scheel

A dissertation submitted in partial fulfillment of the requirements for the Doctor of Philosophy Degree State University of New York College of Environmental Science and Forestry Syracuse, New York April 2020

Department of Chemistry

Approved by:

Christopher T. Nomura, Major Professor John C. Stella, Chair, Examining Committee Avik P. Chatterjee, Department Chair S. Scott Shannon, Dean, The Graduate School

Acknowledgements

The ESF community has played a pivotal role in shaping me into the person I am today; I have discovered new passions, formed many incredible friendships, and learned so much about the environment from many different angles. There are too many people here who have touched my life to properly thank you all, so to the ESF community at large: thank you for helping me become the person I want to be. May you continue to inspire future generations!

The completion of this dissertation would not have been possible without the support of my committee members, Dr. Chris Nomura, Dr. Art Stipanovic, and Dr. Lee Newman. Thank you for your insight and guidance, as well as your patience throughout the research process. I am especially grateful to my advisor Dr. Chris Nomura. You have taught me what it means to be a good scientist. I would also like to extend my sincere thanks to Dr. Erica Majumder. Although our time together at ESF was brief, I truly value the scientific, professional, and personal insight you offered me in the final year of my research. I look forward to seeing the great things you will accomplish here at ESF.

To my colleagues in the Nomura Research Group, thank you for both your friendship and your endless assistance in the lab. I’m extremely grateful to Dr. Ben Lundgren, Dr. Zaara Sarwar, and Dr. Ata Pinto; thank you for devoting your time to training me in all manner of lab techniques, and for the research opportunities within your own projects that you offered me. I would also like to thank all the undergraduates who worked with me on my research. You allowed me to become a better mentor, and your work in the lab was instrumental to this research. I am particularly indebted to Sarah Alam el Din and Seamus McKenney, whose work laid the foundations for the research on nanoparticles and high-density fermentation, respectively.

And finally, to my family. To my mother, Carla: for always teaching and supporting me, especially when I didn’t appreciate it as you deserved. To my grandmother, Marcia: for instilling scientific curiosity in me for as long as I can remember. To my partner, Megan Gallagher: for your encouragement and your inspiring work ethic, which impresses me to this day. And to my best friend, Ryan Crandall: for your enduring friendship, your endless enthusiasm, and for teaching me nearly everything I know about bugs.

ii

Table of Contents List of Tables ...... vi List of Figures ...... vii List of Appendices ...... ix Abstract ...... xii Chapter 1: Introduction and literature review for polyhydroxyalkanoates ...... 1 1.1 Introduction to polyhydroxyalkanoates ...... 1 1.2 PHA composition and physical properties ...... 2 1.3 PHA biosynthesis in native producers ...... 7 1.4 PHA biosynthesis in Escherichia coli LSBJ ...... 9 1.4.1 Engineering of E. coli LS5218...... 9 1.4.2 SCL and MCL biosynthesis pathways ...... 10 1.4.3 Transcriptional regulation of MCL PHA biosynthesis pathway...... 13 1.5 Recent developments in the biosynthesis of chemically modifiable PHAs ...... 28 1.5.1 Fluorination of the PHA α-carbon ...... 29 1.5.2 Development and biosynthesis of azido-PHAs...... 35 1.5.3 Biosynthesis and modifications of unsaturated PHAs ...... 40 1.6 Objectives of this dissertation ...... 42 1.7 References ...... 44 Chapter 2: Enhancing poly(3-hydroxyalkanoate) production in Escherichia coli by the removal of the regulatory gene arcA...... 70 2.1 Abstract ...... 70 2.2 Introduction ...... 71 2.3 Methods ...... 75 2.3.1 Materials ...... 75 2.3.2 Gene deletions ...... 76 2.3.3 PHA production ...... 77 2.3.4 GC analysis ...... 78 2.3.5 Polymer purification and molecular weight determination ...... 78 2.3.6 Growth profiles of LSBJ and RSC02 ...... 79 2.3.7 RNA extraction and qPCR analysis ...... 80 2.4 Results ...... 81 2.4.1 PHA homopolymer production ...... 81

iii

2.4.2 Molecular weight comparison ...... 84 2.4.3 Growth profile of LSBJ and RSC02 ...... 85 2.4.4 PHA yield at 24 vs 48 hours ...... 86 2.4.5 Relative gene expression of RSC02 ...... 87 2.5 Discussion ...... 88 2.6 References ...... 94 Chapter 3: Increased production of the value-added biopolymers poly(R-3-hydroxyalkanoate) and poly(γ-glutamic acid) from hydrolyzed paper recycling waste fines ...... 104 3.1 Abstract ...... 104 3.2 Introduction ...... 105 3.3 Methods ...... 106 3.3.1 Hydrolysate from waste fines ...... 106 3.3.2 Media and cultivation ...... 107 3.3.3 Analytical procedures ...... 109 3.4 Results ...... 110 3.4.1 Hydrolysate characterization ...... 110 3.4.2 PHA analysis ...... 110 3.4.3 PGA analysis ...... 113 3.5 Discussion ...... 113 3.6 References ...... 117 Chapter 4: Optimizing a fed-batch high-density fermentation process for medium chain-length poly(3-hydroxyalkanoates) in Escherichia coli ...... 126 4.1 Abstract ...... 126 4.2 Introduction ...... 127 4.3 Methods ...... 131 4.3.1 Media and inoculum preparation ...... 131 4.3.2 E. coli engineering and plasmid construction ...... 133 4.3.3 Shake flask production ...... 135 4.3.4 Low-density fed-batch production ...... 135 4.3.5 High-density fed-batch production ...... 137 4.3.6 PHA quantification ...... 139 4.3.7 10-azidodecanoic acid synthesis ...... 140 4.3.8 PHA purification and molecular weight analysis ...... 140

iv

4.3.9 Calculations ...... 142 4.4 Results ...... 143 4.4.1 Preliminary trials and the effect of pH ...... 143 4.4.2 Optimization of PHD production ...... 146 4.4.3 Production of alternative MCL PHA ...... 149 4.5 Discussion ...... 152 4.6 References ...... 159 Chapter 5: Preliminary data for a poly(3-hydroxyoctanoate-co-3-hydroxy-10-azidodecanoate) nanoparticle drug delivery system ...... 168 5.1 Abstract ...... 168 5.2 Introduction ...... 169 5.3 Methods ...... 172 5.3.1 Bulk polymer biosynthesis ...... 172 5.3.2 Endotoxin removal and analysis ...... 173 5.3.3 Molecular weight determination ...... 174 5.3.4 Production of PHA nanoparticles ...... 175 5.3.5 Quantification of paclitaxel by HPLC ...... 176

5.3.6 Conjugation of PEG to PHODN3 and PLGA nanoparticles ...... 177 5.4 Results ...... 177 5.4.1 Endotoxin removal and molecular weight analysis ...... 177 5.4.2 Comparison of nanoparticle production methods ...... 180 5.4.3 Paclitaxel encapsulation efficiency ...... 182 5.5 Discussion ...... 183 5.6 References ...... 186 Chapter 6: Conclusions and future outlook ...... 193 6.1 Discussion ...... 193 6.2 References ...... 195

v

List of Tables

Table 1.1: Physical properties of some representative PHAs and conventional plastics...... 6 Table 2.1: Regulation Targets of ArcA and OmpR ...... 75 Table 2.2: Strains, plasmids, and primers...... 76 Table 2.3: PHA yield comparison between LSBJ and RSC02...... 83 Table 2.4: PHA molecular weight data...... 84 Table 3.1: Bacterial strains and plasmids...... 107 Table 4.1: Strains, plasmids, and primers...... 132 Table 4.2: Summary of Bioreactor Conditions ...... 138 Table 4.3: PHD Molecular Weight from Optimization Experiments ...... 149 Table 4.4: Alternative MCL PHA Molecular Weights ...... 152

vi

List of Figures

Figure 1.1: Chemical structures of common representative polyhydroxyalkanoates (PHAs) ...... 3 Figure 1.2: Examples of uncommon and functional PHAs ...... 4 Figure 1.3: Examples of SCL PHA biosynthesis pathways in native producers ...... 8 Figure 1.4: De novo biosynthesis of PHB from acetyl-CoA by recombinant E. coli LSBJ ...... 11 Figure 1.5: Biosynthesis of PHA in E. coli LSBJ utilizing short-chain-length (SCL) and medium-chain-length (MCL) fatty acids ...... 13 Figure 1.6 FadR domain map ...... 15 Figure 1.7: Mechanism of transcriptional regulation by FadR and FadR601(L55Q) ...... 16 Figure 1.8: ArcA domain map ...... 19 Figure 1.9: ArcB domain map ...... 20 Figure 1.10: Simplified, proposed model of the activation of ArcA by ArcB ...... 22 Figure 1.11: Summary of ArcA transcriptional regulation ...... 26 Figure 1.12: Recombinant pathway for 2-fluoro-PHA synthesis in Escherichia coli ...... 31 Figure 1.13: Recent methods for base-mediated defluorosilylation ...... 34 Figure 1.14: Examples of how click reactions can be used to expand functionality of PHA polymers ...... 37 Figure 1.15: Examples of alkynyl-PHA and bromo-PHA biosynthesis by Pseudomonas oleovorans...... 39 Figure 1.16: Cross-linked poly(3-hydroxybutyrate-co-3-hydroxy-10-undecenoate) (PHBUe) produced via thiol-ene click chemistry...... 41 Figure 2.1: Biosynthesis of PHA in E. coli LSBJ utilizing short-chain-length (SCL) and medium-chain-length (MCL) fatty acids ...... 73 Figure 2.2: Comparison of poly(3-hydroxydecanoate) (PHD) yield between LSBJ and mutant strains RSC02, RSC04, and RSC06 ...... 82 Figure 2.3: Growth profile of E. coli LSBJ and RSC02 during PHA-biosynthesis utilizing sodium butyrate (C4) or decanoic acid (C10) ...... 86 Figure 2.4: Comparison of PHA yield as a percentage of dry weight between LSBJ and RSC02 at 24 h and 48 h, utilizing either sodium butyrate (C4) or decanoic acid (C10) ...... 87 Figure 2.5: Relative gene expression of RSC02 compared to LSBJ, normalized to rpoD and measured as fold changes...... 88

vii

Figure 3.1: PHA produced by E. coli strains BW25113 and LSBJ harboring pBBRSTQKAB or pBBRSTQKAB and pTrcGK ...... 111 Figure 3.2: 3HB monomer content of PHB-co-MCL produced by E. coli strains BW25113 and LSBJ harboring pBBRSTQKAB and pTrcGK ...... 112 Figure 3.3: MCL monomer content of PHB-co-MCL produced by E. coli strains BW25113 and LSBJ harboring pBBRSTQKAB and pTrcGK ...... 113 Figure 4.1: Simultaneous glucose consumption and PHA biosynthesis in E. coli LSBJ ...... 130 Figure 4.2: Results of preliminary fermentations that did not result in polymer accumulation 145 Figure 4.3: Optimization of PHD biosynthesis ...... 148 Figure 4.4: Alternative MCL PHA biosynthesis ...... 151 Figure 5.1: Proposed synthesis of targeted nanoparticles for drug delivery ...... 171

Figure 5.2: Endotoxin present in crude PHBDN3 and PHODN3, and average molecular weight

(Mn) of crude polymer samples as determined by GPC ...... 178 -1 Figure 5.3: Change in endotoxin content (EU mg ) and molecular weight (Mn) over successive methanol (MeOH) purifications for PHBDN3 (A, C) and PHODN3 (B, D)...... 179

Figure 5.4: Survey of PHODN3 nanoparticle production methods by size ...... 181

Figure 5.5: Size comparison of PHODN3 nanoparticles produced by nanoprecipitation ...... 182 Figure 5.6: Encapsulation efficiency of polymeric nanoparticles ...... 183

viii

List of Appendices

Appendix A: Nuclear Magnetic Resonance Spectroscopy Data ...... 199 Fig. A1: 1H-NMR (600 MHz, CDCl3) spectrum of poly(R-3-hydroxybutyrate) (PHB) extracted from E. coli RSC02 ...... 199 Fig. A.2: 1H-NMR (600 MHz, CDCl3) spectrum of poly(R-3-hydroxyvalerate) extracted from E. coli RSC02 ...... 200 Fig. A3: 1H-NMR (600 MHz, CDCl3) spectrum of poly(R-3-hydroxyhexanoate) extracted from E. coli RSC02 ...... 200 Fig. A4: 1H-NMR (600 MHz, 50% D2O) of crude hydrolysate in a 1:10 dilution with glucosamine as an internal standard ...... 201 Fig. A5: 1H-NMR (600 MHz, D2O) of poly(γ-glutamic acid) (PGA) ...... 201 Fig. A6: COSY spectra (600 MHz, D2O) of poly(γ-glutamic acid)...... 202 Fig. A7: 1H-NMR (600 MHz, CDCl3) spectrum of poly(R-3-hydroxy-10-undecenoate) extracted from E. coli LSBJ after bioreactor fermentation ...... 203 Fig. A8: 1H-NMR (600 MHz, CDCl3) spectrum of poly(R-3-hydroxy-10-bromodecanoate) extracted from E. coli LSBJ after bioreactor fermentation ...... 203 Fig. A9: 1H-NMR (800 MHz, CDCl3) spectrum of poly(R-3-hydroxyoctanoate-co-3-hydroxy- 10-azidodecanoate) extracted from E. coli LSBJ after 24 h bioreactor fermentation ...... 204 Fig. A10: 1H-NMR (800 MHz, CDCl3) spectrum of poly(R-3-hydroxyoctanoate-co-3- hydroxy-10-azidodecanoate) extracted from E. coli LSBJ after 48 h bioreactor fermentation ...... 204 Fig. A11: 1H-NMR (600 MHz, DMSO-d6) spectrum of poly(R-3-hydroxyoctanoate-co-3- hydroxy-10-azidodecanoate)-graft-dibenzocyclooctyne-poly(ethylene glycol methyl ether) 205 References ...... 205 Appendix B: Nanoparticle Methods Development ...... 206 Nanoparticle production by single emulsion ...... 206 Nanoparticle production by nanoprecipitation ...... 206 References ...... 207 Appendix C: Gel permeation chromatography data ...... 209 Fig. C1: GPC chromatogram of purified poly(3-hydroxybutyrate) (C4) produced by RSC02 ...... 209 Fig. C2: GPC chromatogram of purified poly(3-hydroxyvalerate) (C5) produced by LSBJ (A) and RSC02 (B) ...... 210 Fig. C3: GPC chromatogram of purified poly(3-hydroxyhexanoate) (C6) produced by LSBJ (A) and RSC02 (B) ...... 210 Fig. C4: GPC chromatogram of purified poly(3-hydroxyheptanoate) (C7) produced by LSBJ (A) and RSC02 (B) ...... 211

ix

Fig. C5: GPC chromatogram of purified poly(3-hydroxyoctanoate) (C8) produced by LSBJ (A) and RSC02 (B) ...... 211 Fig. C6: GPC chromatogram of purified poly(3-hydroxydecanoate) (C10) produced by LSBJ (A) and RSC02 (B) ...... 212 Fig. C7: GPC chromatogram of purified poly(3-hydroxydodecanoate) (C12) produced by LSBJ (A) and RSC02 (B) ...... 212 Fig. C8: GPC chromatogram of purified poly(3-hydroxydecanoate) produced by LSBJ via fed- batch fermentation (Trial 9A) ...... 213 Fig. C9: GPC chromatogram of purified poly(3-hydroxydecanoate) produced by RSC02 via fed-batch fermentation (Trial 10B) ...... 213 Fig. C10: GPC chromatogram of purified poly(3-hydroxydecanoate) produced by LSBJ via fed-batch fermentation (Trial 12A) ...... 214 Fig. C11: GPC chromatogram of purified poly(3-hydroxydecanoate) produced by LSBJ via fed-batch fermentation (Trial 13A)...... 214 Fig. C12: GPC chromatogram of purified poly(3-hydroxydecanoate) produced by LSBJ via fed-batch fermentation (Trial 15A)...... 215 Fig. C13: GPC chromatogram of purified poly(3-hydroxydecanoate) produced by LSBJ via fed-batch fermentation (Trial 16A) ...... 215 Fig. C14: GPC chromatogram of purified poly(3-hydroxydecanoate) produced by LSBJ via fed-batch fermentation (Trial 17A) ...... 216 Fig. C15: GPC chromatogram of purified poly(3-hydroxydecanoate) produced by LSBJ via fed-batch fermentation (Trial 18A) ...... 216 Fig. C16: GPC chromatogram of purified poly(3-hydroxyhexanoate) produced by LSBJ via fed-batch fermentation (Trial 19A) ...... 217 Fig. C17: GPC chromatogram of purified poly(3-hydroxyoctanoate) produced by LSBJ via fed-batch fermentation (Trial 20A)...... 217 Fig. C18: GPC chromatogram of purified poly(3-hydroxy-10-undecenoate) produced by LSBJ via fed-batch fermentation (Trial 21A)...... 218 Fig. C19: GPC chromatogram of purified poly(3-hydroxy-10-bromodecanoate) produced by LSBJ via fed-batch fermentation (Trial 22A) ...... 218 References ...... 219 Appendix D: Bioreactor Run Parameters ...... 220 Fig. D1: Trial 9A...... 220 Fig. D2: Trial 9B...... 221 Fig. D3: Trial 10A...... 222 Fig. D4: Trial 11A...... 223 Fig. D5: Trial 12A...... 224 Fig. D6: Trial 13A...... 225

x

Fig. D7: Trial 14...... 226 Fig. D8: Trial 15A...... 227 Fig. D9: Trial 16A...... 228 Fig. D10: Trial 17A...... 229 Fig. D11: Trial 18...... 230 Fig. D12: Trial 19...... 231 Fig. D13: Trial 20...... 232 Fig. D14: Trial 21...... 233 Fig. D15: Trial 22...... 234 Fig. D16: Trial 23...... 235 Fig. D17: Trial 24...... 236 Appendix E: HPLC Sample Chromatograms ...... 237 Fig. E2: Sample HPLC chromatogram of paclitaxel encapsulated in PHODN3 nanoparticle 238 Appendix F: Additional Supplemental Data ...... 239 Table F1: Poly(3-hydroxydecanoate) yield comparison between LSBJ, RSC02, RSC04, and RSC06 ...... 239 Fig. F1: Sample GC chromatogram of purified PHDBr ...... 239 Fig. F2: Sample GC chromatogram of control lyophilized cells ...... 240 Fig. F3: Sample GC chromatogram of methanolyzed cells from Trial 22 at 48h ...... 240 Fig. F4: Calibration curve generated for quantification of methanolyzed PHDBr...... 241 Fig. F5: Schematic of the New Brunswick BioFlo 310 2L bioreactor vessel used in Chapter 4 ...... 241 Appendix G: Peer-reviewed publication covers ...... Error! Bookmark not defined.

xi

Abstract

R. A. Scheel. Enhancing Polyhydroxyalkanoate Biosynthesis in Escherichia coli: A Genetic Engineering and Process Optimization Approach Towards Functionalized Polymeric Nanomedicine, 264 pages, 10 tables (1 in Appendix F), 34 figures (53 in Appendices), 2020. Harvard (author-date) citation style guide was used, with minor formatting changes for publication in Frontiers in Bioengineering and Biotechnology.

Recombinant Escherichia coli is a desirable bacterial chassis for the biosynthesis of many biological compounds including poly(3-hydroxyalkanoates) (PHAs), a class of naturally occurring biodegradable polyesters with promising biomedical and material applications. PHAs are very structurally diverse, and control over their composition allows the production of unique materials with tunable physical and chemical properties. The major limitation to widespread use of PHAs is their high cost and low volumetric productivity, particularly for a desirable class of PHAs with medium chain-length (MCL) side chains. In this work, transcriptional regulators of fatty acid degradation were removed from E. coli in an effort to enhance MCL PHA biosynthesis. Deletion of arcA was successful in improving yields due to increased expression of fadD and fadL in the mutant strains. Although modest improvements in yield were observed in this mutant, a fed-batch high-density fermentation process was developed to further enhance MCL PHA yields using an industrially relevant bioprocess. Through process optimization, some of the highest yields and productivities were achieved to date, including for PHAs with chemically modifiable functional groups. One of these novel functional polymers, a copolymer of 3-hydroxyoctanoate and 3- hydroxy-10-azidodecanoate, was investigated as a targeted nanoparticle drug delivery system. Preliminary results show high encapsulation efficiency of paclitaxel and a surface chemistry amenable to rapid and efficient Click chemistry. In addition to work with MCL PHAs, short chain- length PHAs from a lignocellulosic waste stream, hydrolyzed paper mill waste fines. This work in its entirety demonstrates several successful strategies for improving the availability of PHAs with controlled compositions.

Keywords: poly[(R)-3-hydroxyalkanoates], high-density fermentation, metabolic engineering, biopolymer, targeted drug delivery, nanoparticles

R. A. Scheel Candidate for the degree of Doctor of Philosophy, April 2020 Christopher T. Nomura, Ph.D. Department of Chemistry State University of New York College of Environmental Science and Forestry, Syracuse, New York

xii

Chapter 1: Introduction and literature review for polyhydroxyalkanoates

1.1 Introduction to polyhydroxyalkanoates

Petroleum-based plastics have been rapidly accumulating in the environment for the last

50 years, a process which has only been accelerating. An estimated 250,000 tons of plastic contaminates the world’s oceans, posing a significant threat to wildlife that we are unable to remediate, and that may ultimately end up in our own food supply (Barboza et al., 2018; Eriksen et al., 2014). This has led to growing concerns for the health of the environment, and there is increasing demand for sustainable plastic alternatives (Lazarevic et al., 2010).

Poly-[(R)-3-hydroxyalkanoates] (PHAs) comprise a large class of natural polyesters originally found in . In native PHA-producing microorganisms, these materials are biosynthesized from intracellular fatty acid derivatives or central metabolites and accumulated as a response to increased levels of carbon rich nutrients while experiencing a shortage of other essential elements in the environment (e.g. N, P, O, etc.) (Lu et al., 2009). These biopolymers have enjoyed increasing interest worldwide and have influenced and driven studies in a multitude of disciplines such as microbiology, metabolic engineering, biochemistry, and materials chemistry.

Much of the research to date has focused on the organisms, pathways, and enzymes used to produce

PHA materials.

The predominant class of PHA materials are characterized by a polyester backbone product of the biosynthetic polycondensation of (R)-3-hydroxyacyl-CoA thioesters. These polymers and their bulk physical properties are differentiated and classified by the length of their hydrocarbon sidechains. Whereas PHAs with sidechains containing 0-2 carbon atoms are referred to as short- chain-length (SCL) PHAs, those with sidechains containing 3-11 carbon atoms are commonly

1 identified as medium-chain-length (MCL) PHAs. The former display physical properties typically associated with crystalline thermoplastics while the latter generally possess low crystallinity and elastomeric character (Barham et al., 1992). Poly[(R)-3-hydroxybutyrate] (PHB) is the most well characterized, being the first PHA discovered and the most commonly produced in nature

(Lemoigne, 1926). However, the physical properties of PHB are relatively unattractive, producing a stiff and brittle material (Barham et al., 1992). Copolymerization of PHB and with other monomers such as MCL 3-hydroxyacyl species can significantly increase the flexibility and toughness, and current research is increasingly dominated by these endeavors (Hassan et al., 2006;

Noda et al., 2010).

1.2 PHA composition and physical properties

There is great diversity in the structure of PHAs, which is predominantly localized to the side chains. As previously mentioned, these are most often categorized as SCL (0-2 carbons per side chain) and MCL (3-11 carbons per side chain), although long chain-length (LCL) PHAs have been observed with side chains up to 13 carbons (Lee et al., 1995). Most PHAs form an ester linkage via a β-hydroxyl group, but notable exceptions such as poly(4-hydroxybutyrate) do exist

(Hein et al., 1997; Söhling and Gottschalk, 1996; Valentin and Dennis, 1997). A selection of some representative PHAs is shown in Fig. 1.1, which includes the commercially available group of

SCL-co-MCL PHAs known as Nodax™, in addition to some common SCL and MCL PHAs.

2

Figure 1.1: Chemical structures of common representative polyhydroxyalkanoates (PHAs). Abbreviations: P3HB, poly[(R)-3-hydroxybutyrate]; P4HB, poly(4-hydroxybutyrate); PHHx, poly[(R)-3-hydroxyhexanoate]; PHO, poly[(R)-3-hydroxyoctanoate]; PHD, poly[(R)-3- hydroxydecanoate]; PHBV, poly[(R)-3-hydroxybutyrate-co-3-hydroxyvalerate]; PHB-co-4HB, poly[(R)-3-hydroxybutyrate-co-4-hydroxyvalerate]; PHBHHx, poly[(R)-3-hydroxybutyrate-co-3- hydroxyhexanoate].

3

In addition to the common PHAs shown in Fig. 1.1, numerous examples are documented in the literature of uncommon PHAs containing functional groups. These are addressed in greater detail at the end of this Chapter. The composition of PHAs depends largely on the type of microorganism that produces them; Pseudomonas putida and P. oleovorans are two well-studied examples of MCL PHA producers that can incorporate 3-hydroxyacyl monomers of varying lengths and degrees of saturation, while other commonly described organisms such as Cupriavidus necator and Aeromonas caviae biosynthesize PHB and PHBHHx, respectively (Lu et al., 2009).

Through metabolic engineering and the selection of unique carbon substrates, an even greater variety than what is observed in nature can be achieved (Dong et al., 2019; Matsumoto and

Kageyama, 2019; Mizuno et al., 2018; Sparks and Scholz, 2008; Thuronyi et al., 2017). Some examples of these can be found in Fig. 1.2.

Figure 1.2: Examples of uncommon and functional PHAs. Abbreviations: P2M3HB, poly(2- methyl-3-hydroxybutyrate); P2MHV, poly(2-methyl-3-hydroxyvalerate); P2H4MV, poly(2- hydroxy-4-methylvalerate); P2F3HB, poly(2-fluoro-3-hydroxybutyrate); PHUe, poly(3-hydroxy- 10-undecenoate); PHUy, poly(3-hydroxy-10-undecynoate); PHDN3, poly(3-hydroxy-10- azidodecanoate); PHDBr, poly(3-hydroxy-10-bromodecanoate).

4

The physical properties of PHAs are largely determined by their chemical composition.

Homopolymeric PHB is highly crystalline, generally ranging from 60-80% crystallinity depending on the method of extraction and/or post-isolation processing (Chanprateep, 2010; Doi and Abe,

1990; Tappel et al., 2012b). In addition, PHB has a high Young’s modulus similar to the petrochemically-derived plastic polypropylene (Werber and Baptist, 1964). However, PHB has considerably lower toughness and elongation to break than polypropylene, rendering it quite brittle

(Chanprateep, 2010; Sudesh et al., 2000). In comparison, several other PHA homopolymers exhibit elastomeric properties and a greater degree of toughness. Poly(4-hydroxybutyrate), a PHA first isolated from the anaerobic bacterium Clostridium kluyverii, has an elongation to break of 1000%, and the copolymerization of 3HB with 4HB monomers confers an increasing degree of toughness to the resulting material as the monomer % of 4HB increases (Doi and Abe, 1990; Saito and Doi,

1994). Copolymers of 3HB and MCL monomer units also have attractive physical properties, with increasing MCL monomer content conferring increasingly elastomeric qualities, although these are relatively uncommon in nature (Noda et al., 2005). MCL copolymers, which are most often ter- and quadripolymers of various MCL monomers, are amorphous and characterized by low or indiscernible melting points (Gagnon et al., 1992). A summary of the physical properties of some common PHAs and conventional plastics is listed in Table 1.1.

5

Table 1.1: Physical properties of some representative PHAs and conventional plastics.

Crystallinity Young’s Tensile Elongation to Tg (°C) Tc (°C) Tm (°C) References (%) Modulus Strength Break (%) (MPa) (MPa) PHB 60 to 80 3500 40 0.4-45 -1.2 to 45 to 177 (Chanprateep, 2010; Doi and 10 95 Abe, 1990; Tappel et al., 2012b) PHBV (%3HV) 9 58 1900 37 6 125 (Doi and Abe, 1990) 17.2 55 ~775 ~22 ~25 172 (Gogotov et al., 2010) 22.2 47 ~450 ~14 ~10 166 (Gogotov et al., 2010) P4HB 34 70 to 149 50 to 104 1000 -51 to (Martin and Williams, 2003; -46 Saito and Doi, 1994) PHBHHx 7 289 17.3 23.6 125/135 (Asrar et al., 2002) 9.5 155 8.8 43.0 (Asrar et al., 2002) MCL PHAs (dominant monomers) 3HO,3HHx -37 59 (Preusting et al., 1990) 3HOU,3HO,3HHU,3HHx -37 (Preusting et al., 1990) 3HD,3HO,3HHx 12 to 15 7 to 10 300 to 450 -35 55 (Gagnon et al., 1992) 3HO,3HB 12 to 170 1.7 to 10.5 53 to 230 -26 to - 158 (Tappel et al., 2012a) 1.4 Conventional Plastics LDPE 200 10 620 -30 130 (Thompson et al., 2004) PP 50 to 70 1700 38 400 -10 176 (Thompson et al., 2004) PETE 30 to 50 2900 70 1000 69 267 (Doi and Abe, 1990) PS 76.8 0.36 (Chanprateep, 2010) Blank cells indicate the value was not determined from the referenced study. Tg (glass-transition temperature), Tc (crystallization temperature by differential scanning calorimetry, DSC), Tm (melting temperature). Polymer abbreviations; PHB, poly[(R)-3- hydroxybutyrate; PHBV, poly[(R)-3-hydroxybutyrate-co-3-hydroxyvalerate]; P4HB, poly(4-hydroxybutyrate); PHBHHx, poly[(R)-3- hydroxybutyrate-co-3-hydroxyhexanoate]; 3HO, 3-hydroxyoctanoate; 3HHx, 3-hydroxyhexanoate; 3HOU, 3-hydroxy-7-octenoate; 3HHU, 3-hydroxy-5-hexenoate; 3HD, 3-hydroxydecanoate; 3HB, 3-hydroxybutyrate; LPDE, low-density poly(ethylene); PP, poly(propylene); PETE, poly(ethylene terephthalate); PS, poly(styrene).

6

1.3 PHA biosynthesis in native producers

PHA biosynthesis pathways in native organisms generally fall into one of two categories,

SCL biosynthesis and MCL biosynthesis. PHB, the most common SCL PHA, is produced by a wide variety of organisms using the same set of three enzymes; (1) a β-ketothiolase (PhaA) that catalyzes the condensation of two molecules of acetyl-CoA to form acetoacetyl-CoA, (2) an acetoacetyl-CoA reductase (PhaB) that catalyzes the reduction of acetoacetyl-CoA to form (R)-3- hydroxybutyryl-CoA, and (3) a PHA synthase (PhaC) which catalyzes the formation of an ester bond between two (R)-3-hydroxybutyryl-CoA molecules and begins the elongation of a growing polymer chain (Lu et al., 2009).

PhaC can often accept other hydroxyacyl-CoA substrates, resulting in the formation of

PHV, P4HB, or the copolymers PHBV and P3HB4HB. PHV is formed from the metabolism of propionate, which can be supplied exogenously or formed from oxaloacetate in a reversible reaction of gluconeogenesis (Robbins, 1988). Propionate is then converted into propionyl-CoA and condensed with a molecule of acetyl-CoA by an alternative β-ketothiolase (BktB) to form β- ketovaleryl-CoA (Slater et al., 1998). This intermediate is then reduced by PhaB and subsequently polymerized by PhaC, resulting in PHV or PHBV.

P4HB is produced by Clostridium kluyverii from succinyl-CoA, which is converted into succinate semialdehyde by a succinate semialdehyde dehydrogenase (SucD) (Steinbüchel and

Valentin, 1995). This intermediate is reduced to 4HB by 4-hydroxybutyrate dehydrogenase

(4hbD), converted to 4-hydroxybutyryl-CoA by a succinyl-CoA:CoA transferase (OrfZ, aka Cat1), and is polymerized by PhaC. These pathways are shown in Fig. 1.3.

7

Figure 1.3: Examples of SCL PHA biosynthesis pathways in native producers. (A) Production of PHB from sugars (acetyl-CoA). (B) Production of PHV from oxaloacetate. (C) Biosynthesis of P4HB from succinyl-CoA.

8

1.4 PHA biosynthesis in Escherichia coli LSBJ

In order to address the difficulties observed with controlling PHA composition, Tappel et al. engineered a recombinant E. coli strain that was able to exhibit precise control in both SCL and

MCL monomer incorporation from fatty acid substrates (2012b). E. coli was chosen as a production chassis for several reasons: its genome is very well understood and there are simple methods described to manipulate it, it can grow rapidly to high density and tolerate a wide variety of growth conditions and media, and it cannot naturally produce PHA which allows for greater control over the PHA biosynthesis pathway. The construction of this strain is briefly described below.

1.4.1 Engineering of E. coli LS5218

Since control of repeating unit composition is critical to produce polymers with desirable, reproducible properties, the Nomura lab engineered E. coli to produce PHA polymers with defined repeating unit compositions from fatty acids (Tappel et al., 2012a). The parental strain used was

E. coli LS5218 (Spratt et al., 1981), which carries two significant mutations that make it ideal for producing PHA polymers from fatty acids: (1) a fadR mutation for enhanced expression of genes encoding enzymes involved in β-oxidation, described in more detail below; and (2) an atoC(Con) mutation which results in the constitutive expression of the Ato short-chain-fatty acid uptake enzymes in E. coli. This strain was further engineered to control the production of enoyl-CoA by deleting the fadB (Yang et al., 1995) and fadJ (Campbell et al., 2003) genes (E. coli LSBJ), both of which can catalyze the (S)-specific hydration of enoyl-CoA, resulting in a roadblock in β- oxidation (Figure 1.5). This resulted in a strain where any exogenous fatty acid supplied was pooled as its enoyl-CoA intermediate with no loss of carbon from the substrate. For example, supplementation with a 4-carbon fatty acid (butyrate) results in a butenoyl-CoA intermediate, and

9 supplementation with an 8-carbon fatty acid (octanoate) results in pooling of an octenoyl-CoA intermediate. Thus, specific fatty acid feeding regimens, coupled with expression of the (R)- specific enoyl-CoA hydratase (PhaJ4) and PhaC1(STQK) engineered synthase, results in production of PHA polymers with defined repeating unit compositions (Takase et al., 2003; Tappel et al., 2012b).

1.4.2 SCL and MCL biosynthesis pathways

SCL PHA biosynthesis can be accomplished in E. coli LSBJ from glycolytically-derived acetyl-CoA via the heterologous expression of three enzymes: a β-ketothiolase (PhaA), an acetoacetyl-CoA reductase (PhaB), and a PHA synthase (PhaC). This process is functionally the same as that described above for native producers (Fig. 1.4).

10

Figure 1.4: De novo biosynthesis of PHB from acetyl-CoA by recombinant E. coli LSBJ.

The biosynthesis of PHA in E. coli LSBJ utilizing short-chain-length (SCL) and medium- chain-length (MCL) fatty acids is shown in Fig. 1.5. The absence of fadB and fadJ in E. coli LSBJ in combination with the plasmid-borne recombinant enzymes PhaC1(STQK) and PhaJ4 establishes a linear pathway for the production of PHA polymers from free fatty acids.

Extracellular fatty acids are transported across the outer membrane dependent on size; SCL and shorter MCL fatty acids can diffuse across the outer membrane, while longer MCL fatty acids (12 carbons and higher) can be transported by the long-chain fatty acid transporter FadL (Lepore et

11 al., 2011). Inner membrane transport and activation is accomplished by the SCL-specific Ato system (AtoEAD) or the MCL-specific acyl-CoA synthetase FadD (Kameda and Nunn, 1981;

Theodorou et al., 2006). Acyl-CoA substrates are converted into enoyl-CoA by the acyl-CoA dehydrogenase enzyme FadE (Campbell and Cronan, 2002), and are unable to proceed further through β-oxidation due to the absence of FadB and FadJ. The enoyl-CoA pool is then converted to (R)-3-hydroxyacyl-CoA by the R-specific enoyl-CoA hydratase PhaJ4 (Tsuge et al., 2003), and finally polymerized by the PHA synthase PhaC1(STQK) (Takase et al., 2004, 2003). This system allows for the biosynthesis of PHA polymers with tightly controlled repeating unit composition, as the number of carbons present in the fatty acid substrate is retained as the total number of carbons in each repeating unit. This was previously demonstrated by Tappel et al. with both single fatty acid substrates, resulting in PHA homopolymers with the same number of carbon atoms as the fatty acid precursors, as well as dual fatty acid co-substrates to produce PHA copolymers with controlled composition (2012a, 2012b).

12

Figure 1.5: Biosynthesis of PHA in E. coli LSBJ utilizing short-chain-length (SCL) and medium- chain-length (MCL) fatty acids.

1.4.3 Transcriptional regulation of MCL PHA biosynthesis pathway

Due to the inability of E. coli to natively biosynthesize PHA, regulation of MCL PHA biosynthesis primarily occurs at the level of fatty acid uptake and catabolism. These processes are tightly regulated by a number of transcription factors, which will be described here.

13

1.4.3.1 Regulation by FadR

One of the primary regulators of fatty acid metabolism is the transcriptional dual regulator

FadR, or the fatty acid degradation regulon (Overath et al., 1969). FadR is responsible for repressing genes necessary for fatty acid degradation (β-oxidation) and activating genes necessary for fatty acid biosynthesis to control carbon metabolism in response to exogenous fatty acids. FadR interacts with the upstream region of genes within its regulon by binding to the palindromic consensus binding sequence -TGGNNNNNCCA- (Xu et al., 2001). When this consensus sequence overlaps the -35 or -10 region (and the region immediately upstream of the -10), as is the case for fad genes, FadR binding prevents the recruitment of RNA polymerase (RNAP) and inhibits transcription (Black and DiRusso, 1994; Cronan and Subrahmanyam, 1998; Henry and Cronan,

1992). If the FadR binding site lie upstream of the -35 region, as it does for the fab genes (fatty acid biosynthesis), then the binding of FadR activates transcription by enhancing RNAP recruitment.

Crystal structures have been determined for FadR, the FadR-acyl-CoA complex, and FadR bound to DNA, which have helped to develop a mechanistic understanding of this enzyme (van

Aalten et al., 2000; Xu et al., 2001). The FadR regulator functions as a homodimer, and binds to

DNA through the interaction of the two DNA-Binding domains of the apoenzyme (Fig. 1.6)

(Raman et al., 1997). The DNA-binding domain, which has been determined to be within the first

66 amino acid residues of the N-terminus (Raman et al., 1997), forms a winged helix motif common in prokaryotic transcriptional regulators and is highly conserved in fadR homologs of other prokaryotes (Brennan, 1993; Iram and Cronan, 2005). A large region of the C-terminus forms the dimerization and acyl-CoA binding domains; although the exact residues involved in dimerization are currently unknown, three residues necessary for acyl-CoA binding (G216, S219,

14

W223) have been identified in E. coli MG1655 (DiRusso et al., 1998). This region is much less conserved than the DNA binding domain (~25% homology), and is responsible for significant differences in the preferred fatty acid substrates among bacteria (Iram and Cronan, 2005). In E. coli LS5218, the parental strain of LSBJ, a single nucleotide polymorphism (SNP, L55Q) in the

DNA binding domain prevents DNA binding and relieves fad repression by FadR (Rand et al.,

2017). This mutation, also known as fadR601, is believed to disrupt a hydrophobic interaction among three neighboring residues based on evidence from other E. coli fadR mutants (Zhang et al., 2014).

Figure 1.6 FadR domain map. N-Terminal DNA binding domain contains a SNP in E. coli LS5218 and its derivatives that results in a nonfunctioning FadR. This domain is separated from the C- terminal dimerization / acyl-CoA binding domain by a short linker. Acyl-CoA binding requires three residues in E. coli MG1655 and LS5218, G216, S219, and W223.

The FadR sensing mechanism hinges on the binding of fatty acyl-CoA molecules to the bound apo-enzyme complex (Xu et al., 2001). The binding of a fatty acyl-CoA triggers conformational changes in the protein which releases DNA from the N-terminal DNA-binding domain (Fig. 1.7). This means that the presence of fatty acyl-CoA molecules relieves the repression of β-oxidation genes while simultaneously abolishing FadR-mediated activation of fatty

15 acid biosynthesis. In E. coli MG1655 and many related strains, FadR will only bind long chain- length (LCL) fatty acyl-CoAs, and has the highest affinity towards oleoyl-CoA (Feng and Cronan,

2012; Iram and Cronan, 2006). For this reason, these strains cannot grow on MCL fatty acids such as decanoate due to the inhibition of the β-oxidation pathway. The FadR601 mutation is therefore necessary for the production of MCL PHAs in E. coli LSBJ from fatty acids of 6 to 14 carbons.

Figure 1.7: Mechanism of transcriptional regulation by FadR and FadR601(L55Q). (A) In the absence of acyl-CoA, FadR binds to the promoter region of fad genes and represses transcription. DNA binding is blocked in the nonfunctioning FadR601. (B) In the presence of acyl-CoA, FadR binds acyl-CoA and releases DNA through due to conformational changes. Acyl-CoA binding is not blocked in FadR601, however it has no effect on transcription. This figure was adapted from Feng and Cronan (2012).

1.4.3.2 Regulation by ArcA/ArcB

Another important regulator of fatty acid catabolism is the transcriptional dual regulator

ArcA, which forms the central hypothesis in Chapter 2 and will therefore be reviewed in detail. In

E. coli, ArcA serves as the response regulator of the ArcAB two-component signal transduction

16 system, and plays a pivotal role in controlling the switch from aerobic to anaerobic metabolic processes in response to oxygen availability (Iuchi and Lin, 1988; Lynch and Lin, 1996).

Due to the large number of genes affected by ArcA, researchers initially had difficulty characterizing the protein, and subsequently there are several synonyms present in the literature.

In 1973 and 1976, researchers from Australia published articles in which they identified and named the seg gene (arcA), so named because a strain harboring a mutant gene prevented segregation and replication of the F-factors (Hathaway and Bergquist, 1973; Jamieson and Bergquist, 1976). In

1978, a separate group found that strains with this particular mutation were resistant to male- specific bacteriophages, and later studies in the early 1980’s revealed that these mutants showed inhibition of F-pilus expression, leading to the classification msp and fex, respectively (Buxton et al., 1978; Lerner and Zinder, 1982). Within the same time period, several studies explored the effects of this gene using chemical mutagenesis, producing a set of mutants with varying impacts on sex factor regulation, which led to the new synonym sfrA (Beutin et al., 1981; Beutin and

Achtman, 1979).

The realization that this putative sex factor gene had a larger role in transcriptional regulation did not occur until 1983, when an article was published which made connections between the sex factor regulation studies and a seemingly unrelated paper on the dye gene. Mutants lacking this dye gene were previously shown in 1979 to have increased sensitivity to the dye toluidine blue and several antibiotics (Roeder and Somerville, 1979). Researchers from the UK

(Buxton et al.) found that the putative dye gene occupied the same chromosomal position as the putative fex gene, and through a series of complementation studies determined that dye, fex, and sfrA were all the same gene (Buxton et al., 1983; Buxton and Drury, 1983). They also revealed that the envelope protein profile in Δdye mutants had several differences compared to strains with

17 functional dye protein product, which was believed to be the cause for increased sensitivity to dyes and antibiotics, and the researchers concluded that the dye gene may be involved in the regulation of envelope proteins in addition to sex factor expression (Buxton et al., 1983). This team of researchers continued their exploration of the now-named dye gene, publishing several more articles in which they characterized the Dye protein, analyzed the DNA and amino acid sequences, and found some similarity to the transcriptional regulator OmpR (Buxton and Drury, 1984; Drury and Buxton, 1985).

The final piece of evidence necessary for discovering the true function of the dye(arcA) gene came from an unrelated study published in 1988. Researchers from Harvard Medical School under Dr. Edmund Lin, while studying the transcriptional regulation of aerobic and anaerobic metabolic pathways, created a mutant E. coli strain that retained very high anaerobic expression of many enzymes normally expressed predominantly during aerobic growth (Iuchi and Lin, 1988).

They isolated the gene responsible for this phenotype, which they labelled as arcA (aerobic respiration control), and by mapping the gene to the genome they were able to determine that it occupied the same locus as the previously described dye gene. The group confirmed that the two genes were identical by observing that the arcA mutant was both sensitive to the dye toluidine blue and resistant to the male-specific phage M13 (Iuchi and Lin, 1988). This study became the foundation for the modern understanding of arcA, and its role as a transcriptional dual-regulator.

ArcA belongs to the OmpR family of proteins, which is a typical family of two-component response regulators (Cho et al., 2006; Itou and Tanaka, 2001; West and Stock, 2001). The arcA gene has a nucleotide sequence of 717 bases, encoding a protein of 238 amino acids in length and a molecular weight (MW) of 27,346 Da, as determined computationally and analytically (Buxton

18 and Drury, 1984; Drury and Buxton, 1985). The ArcA protein purifies along with other soluble cytosolic proteins (Buxton and Drury, 1984).

The domain structure for ArcA is relatively simple; as in other OmpR-family proteins, the active enzyme forms a homodimer, and the two domains present include a receiver domain and a

DNA-binding domain (Fig. 1.8) (Cho et al., 2006; Drury and Buxton, 1985). The crystal structure of ArcA has been solved by X-ray crystallography, enhancing our understanding of these domain interactions (Toro-Roman et al., 2005). The receiver domain, also known as the dimerization domain (D1’), is located from amino acids 6-113 (N-terminal) and serves both as the site of dimerization and as the site where the ArcB histidine kinase interacts with and transphosphorylates

ArcA (Drury and Buxton, 1985; Finn et al., 2014; West and Stock, 2001). It is within this domain that the catalytic aspartic acid residue (D54) resides, which activates the enzyme upon its phosphorylation (Drury and Buxton, 1985; Finn et al., 2014). The DNA-binding domain (DB1) is responsible for interacting with the promoter regions of regulons under the control of ArcA, and is located from amino acids 156-232 (C-terminal) (Drury and Buxton, 1985; Finn et al., 2014).

The bound ArcA-P acts as a dual-regulator, capable of inhibiting or activating various genes (Cho et al., 2006).

Figure 1.8: ArcA domain map. N-Terminal dimerization domain contains a catalytic aspartic acid residue (D54), and is separated from the C-terminal DNA-Binding domain by a short linker.

19

For a complete understanding of the function of ArcA, it is also necessary to discuss the structure of ArcB due to its important role in the activation of ArcA. The sensor protein ArcB is classified as a tripartite hybrid histidine kinase, and contains several important domains; two transmembrane domains (TM1 & 2) separated by a short periplasmic domain (P1), a dimerization domain (D1) which also acts as a linker, and primary and secondary transmitter domains (T1 & 2) separated by a receiver domain (R1), shown below in Fig. 1.9 (Ishige et al., 1994; Iuchi and Lin,

1992; Kwon et al., 2000; Nuñez Oreza et al., 2012). The two transmembrane domains serve to anchor the complete enzyme in the inner membrane, but neither they nor the periplasmic domain play an active role in sensing; instead, the cytosolic portion of the enzyme interacts directly with quinones involved in the electron transport chain (Ishige et al., 1994; Iuchi and Lin, 1992; Kwon et al., 2000; Nuñez Oreza et al., 2012). The dimerization domain has the characteristics of a leucine zipper, and forms a homodimer by a coiled-coil motif (Nuñez Oreza et al., 2012). The two transmitter domains along with the receiver all play a role in the phosphorelay system (Nuñez

Oreza et al., 2012; Peña-Sandoval and Georgellis, 2010).

Figure 1.9: ArcB domain map. The N-terminal transmembrane domain consists of two membrane spanning regions (TM1&2) separated by a short periplasmic linker. The dimerization domain (D1) separates the transmembrane domain from the cytosolic phosphorelay domains, which contains the primary and secondary transmitter domains (T1&2) separated by a receiver domain (R1), each with catalytic residues of H292, D576, and H717, respectively.

20

ArcA is activated by phosphorylation under reducing conditions, or conditions of low oxygen, by a phosphorelay system involving the electron transport chain, ArcB, and an aspartic acid residue of ArcA (Iuchi and Lin, 1992; Peña-Sandoval and Georgellis, 2010). Under reducing conditions, the His292 residue of ArcB (T1, Fig. 1.10A) autophosphorylates through interaction with cystosolic ATP( Iuchi and Lin, 1992; Peña-Sandoval and Georgellis, 2010). Through intramolecular reactions, the phosphoryl group is passed to the Asp576 residue (R1, Fig. 1.10A) and finally to the His717 residue (T2, Fig. 1.10A) located near the C-Terminus (Iuchi and Lin,

1992; Peña-Sandoval and Georgellis, 2010). The phosphoryl group is then transferred to the Asp54 residue in the dimerization domain of ArcA (D1’, Fig 1.10A), resulting in the activated form of

ArcA, ArcA-P (Iuchi and Lin, 1992; Peña-Sandoval and Georgellis, 2010). This activated ArcA-

P then proceeds to bind DNA in order to alter gene expression.

Under aerobic conditions, ArcA-P undergoes a reverse phosphorelay and is de-activated.

First, quinones involved in electron transport become oxidized, and silence the kinase function of

ArcB by forming disulfide bonds in the dimerization domain, preventing the phosphorelay in the activating direction (Georgellis et al., 2001, 1998). The phosphorylated ArcA-P proceeds to transfer its phosphoryl group back to the His717 residue of ArcB (T2, Fig. 1.10B), which then transfers it to the receiver domain (R1, Fig. 1.10B) at Asp576 (Georgellis et al., 2001, 1998).

Together, the forward and reverse phosphorelay systems serve to control gene regulation by ArcA in response to the oxygen available in the environment.

21

Figure 1.10: Simplified, proposed model of the activation of ArcA by ArcB. QH2 and Q refer to the reduced and oxidized forms of quinones, respectively. (A) Autophosphorylation of ArcB (The membrane-bound protein) cleaves a phosphoryl group (yellow “P”) from ATP, then transfers this phosphoryl group along the relay to ArcA (orange), which allows DNA binding through allosteric modification. (B) De-phosphorylation of ArcA occurs in aerobic conditions, where quinone- induced disulfide bonds form between ArcB molecules, preventing the cleavage of ATP and forcing the relay in the opposing direction, preventing DNA binding (Georgellis et al., 2001, 1998; Malpica et al., 2004; Nuñez Oreza et al., 2012; Peña-Sandoval and Georgellis, 2010).

As a transcriptional dual regulator, ArcA has the capacity to both repress and activate certain genetic elements. In order to determine which genes ArcA was most likely to regulate, it was necessary for researchers to identify a consensus binding sequence; an extensive 1996 article written by Lynch and Lin used a series of electrophoretic mobility shift assays and restriction enzyme digests to hone in on where ArcA was binding in relation to several model genes, and to

22 determine the consensus binding sequence (Lynch and Lin, 1996). They characterized a putative consensus binding site with the sequence [5’ (A/T)GTTAATTA(A/T) 3’] and used this information to verify that genes known to be regulated by ArcA contained sequences with high similarity to this consensus (Lynch and Lin, 1996).

Once a putative sequence was determined, computational analysis revealed that several genes contained two or three potential ArcA binding sites rather than just one (Lynch and Lin,

1996). These sequences generally appear upstream of the transcription start site in the promoter region, though they can also appear downstream or overlapping the transcription start site as well

(Lynch and Lin, 1996). Since this initial experiment, more in-depth computational analyses have determined that there are as many as 150 potential operons under the direct influence of ArcA, and a revised consensus sequence of [5’ GTTAATTAAATGTT 3’] was proposed and is used as the current consensus for genetic analysis (Keseler et al., 2013; Liu and Wulf, 2004).

Due to the high number of genes and operons affected by ArcA, the overall transcriptional effects are generally reported as large families involved in certain metabolic functions. In anaerobic conditions, when ArcA is phosphorylated and active, several metabolic pathways required for efficient energy utilization in aerobic conditions are repressed (Jiang et al., 2015;

Salmon et al., 2005). These pathways include; the tricarboxylic acid cycle, glycolysis, amino acid transport and catabolism, fatty acid transport and degradation, and quinone oxidation, among many others (Salmon et al., 2005). There are also several metabolic functions that are activated by ArcA under anaerobic conditions, which mainly involved small molecule transport and metabolism, nitrate reduction, and DNA repair (Salmon et al., 2005).

The implications of these changes in cellular metabolism is that cells are able to utilize various energy sources more efficiently in response to oxygen availability. Pathways involving

23 oxygen utilization become down-regulated as the availability of oxygen decreases, and concurrently the expression of pathways needed for alternative electron transport (such as nitrate) and fermentation become up-regulated (Salmon et al., 2005).

Among the many genes regulated by ArcA, there are several transcriptional regulators; this greatly increases the complexity of ArcA regulation through the addition of indirect effects

(Keseler et al., 2013; Salmon et al., 2005). Perhaps the most interesting of these indirect effects is the suppression of the sigma factor RpoS through the down-regulation of the small regulatory

RNA, arcZ (Mandin and Gottesman, 2010; Soper et al., 2010). The small regulatory RNA arcZ acts as a translational activator of rpoS, binding to the rpoS mRNA and recruiting the Hfq chaperone protein, which stabilizes the molecular construct in an “active” conformation where translation can proceed (Soper et al., 2010).

Under anaerobic conditions, when ArcA is active, the expression of arcZ is highly down- regulated, which in turn inhibits the translation of rpoS (Mandin and Gottesman, 2010; Soper et al., 2010). In addition, activated ArcA-P binds to the promoter region of rpoS and inhibits transcription, which further down-regulates rpoS expression (Liu and Wulf, 2004; Mandin and

Gottesman, 2010). Conversely, under aerobic conditions rpoS expression is up-regulated both by the removal of ArcA-P inhibition and by the translational activation of arcZ (Liu and Wulf, 2004;

Mandin and Gottesman, 2010; Soper et al., 2010). However, determining the whole transcriptome effects of this indirect regulation is difficult to discern.

RpoS, or sigma-factor S (σS), is involved in the transcription of a diverse group of genes making up approximately 10% of the E. coli genome (Franchini et al., 2015; Patten et al., 2004;

Rahman et al., 2006). Many of these genes are involved in stress response, and there is a large amount of regulatory overlap between σS and σ70, which have similar consensus sequences, and

24 other stress-response sigma factors (σ54 and σ28) (Dong et al., 2011; Franchini et al., 2015; Patten et al., 2004; Rahman et al., 2006). This complex regulatory network makes it difficult to definitively attribute transcriptional effects to σS. However, studies have found that RpoS (σS) is involved in functions such as stress response, secondary metabolism, nucleic acid synthesis, as well as sugar and polyamine metabolism in response to stress (Keseler et al., 2013; Maciag et al.,

2011). The control of RpoS by ArcA increases the widespread transcriptional effects of the ArcAB two-component signal transduction pathway, and likely assists in the metabolic shift that occurs between anaerobic and aerobic conditions. Fig. 1.11 below displays a summary of the ArcA regulon, which includes aerobic respiration, the tricarboxylic acid cycle, glycolysis, amino acid transport and catabolism, fatty acid transport and degradation, RpoS expression, and quinone oxidation.

25

Figure 1.11: Summary of ArcA transcriptional regulation. Transphosphorylation of ArcA by ArcB leads to the up-regulation of small molecule metabolism and nitrate reduction, the down regulation of aerobic respiration, fatty acid degradation, amino acid catabolism, and RpoS expression, among other pathways (Jiang et al., 2015; Liu and Wulf, 2004; Mandin and Gottesman, 2010; Salmon et al., 2005; Soper et al., 2010). Green arrows refer to activation, and red arrows refer to inhibition.

1.4.3.3 Regulation by CRP-cAMP

The cyclic AMP receptor protein (CRP) – cyclic AMP complex (CRP-cAMP) is another important regulator of fatty acid metabolism in E. coli, responsible for activating transcription of the fad regulon in the absence of a preferred carbon source, glucose (Fic et al., 2009). This is part of larger process known as carbon catabolite repression, which is mediated by a protein phosphorylation cascade via components of the phosphoenolpyruvate (PEP):carbohydrate phosphotransferase system, or PTS (Deutscher, 2008; Franchini et al., 2015). Glucose is transported into the cell by the glucose-specific membrane permease EIICGlc, concomitant with its

26 phosphorylation by the EIIB subunit to form glucose-6-phosphate (Kundig et al., 1964).

Transphosphorylation of EIIB is catalyzed by EIIA to ensure continued glucose transport, which results in an elevated pool of the dephosphorylated EIIA. This in turn reduces the activity of adenylate cyclase (CyaA), which requires phosphorylation to catalyze the synthesis of cAMP. A reduction in available cAMP in reduces the concentration of CRP-cAMP, ultimately down- regulating fad regulon expression. Conversely, when glucose in unavailable the pool of CRP- cAMP in the cytosol is high, activating the transcription of both fadD and fadE (Feng and Cronan,

2012). This has implications for the biosynthesis of MCL PHAs in E. coli LSBJ.

1.4.3.4 Regulation by OmpR/EnvZ

One member of the fad regulon, the LCL fatty acid transporter FadL, is additionally regulated by the OmpR/EnvZ two component response regulator (Aiba and Mizuno, 1990; Forst and Roberts, 1994). Although this regulatory system was implicated in the hypotheses in Chapter

2, it was found to not play a role in MCL PHA biosynthesis in E. coli LSBJ and will therefore only be reviewed briefly.

The OmpR/EnvZ regulatory system is primarily responsible for controlling the expression of outer membrane porins in response to osmolarity, with OmpR serving as the response regulator for this system (Mizuno and Mizushima, 1990). OmpR is a well-characterized transcriptional regulator in E. coli, belonging to a large family of transcription factors with a winged helix DNA binding motif (the OmpR family) (Kato et al., 1995). The outer membrane porins OmpF and

OmpC, both of which form ion-permeable channels and aid in the diffusion of small hydrophilic molecules across the inner membrane, are the primary targets of OmpR regulation (Foo et al.,

2015). Although both porins allow transport of similar molecules, OmpF is characterized by a faster flow rate and is its expression is activated in conditions of low osmolarity. OmpC plays an

27 opposing role; in conditions of high osmolarity ompC expression is activated, and slower flow rates through OmpC help maintain ion homeostasis (Nikaido, 1992; Silhavy and Pratt, 1995).

Regulation by OmpR is controlled by the sensor kinase EnvZ, which autophosphorylates in response to extracellular and intracellular osmolarity and transphosphorylates OmpR, which proceeds to bind to the upstream promoter region of target genes (Forst and Roberts, 1994;

Matsubara and Mizuno, 1999; Wang et al., 2012). Although ompF and ompC are the primary targets for OmpR binding, it has also been shown to modulate the expression of amino acid metabolism, flagellar biosynthesis, and LCL fatty acid transport via the outer membrane porin

FadL (Higashitani et al., 1993; Oshima et al., 2002). Although FadL is sometimes implicated in

MCL PHA biosynthesis, it has been shown to exhibit no activity on fatty acids of 12 carbons or less (Black, 1990).

1.5 Recent developments in the biosynthesis of chemically modifiable PHAs

Significant research effort has centered on expanding the capability of PHA-producing bacteria to incorporate, in a controlled fashion, MCL/SCL and unnatural monomers into copolymers by engineering biosynthetic pathways and expanding the substrate specificities of their enzymatic constituents. There have been several successful examples of this approach, including chimeric PHA synthases with broad substrate specificity (Matsumoto et al., 2009), the development of controllable production platforms such as E. coli LSBJ (Tappel et al., 2012b), and engineered recombinant pathways capable of synthesizing unique copolymers (Li et al., 2017,

2016). These research efforts have yielded slow-yet-steady gains in the production of uncommon

PHA copolymers from a variety of related and unrelated feedstocks.

An underdeveloped area of PHA research, however, has been that of finding solutions to the limited chemical tractability of PHA side chains. Thus far this challenge has been circumvented

28 by applying multistep synthetic routes post-polymerization to adapt them to specific applications and designs. Besides some excellent examples from a few laboratories, meaningful efforts to augment the chemical tractability of PHAs via incorporation of unnatural or uncommon monomers have been sporadic and inadequate. Similarly, organisms such as Pseudomonas oleovorans have been shown to produce exotic and uncommon PHAs biosynthetically, however there has been little progress towards scaling up their production (Lenz et al., 1992).

A great variety of uncommon functional groups have been observed in PHAs, including; terminal alkenes and alkynes (Lageveen et al., 1988; Levine et al., 2015; Tappel et al., 2012b), halogens (Doi and Abe, 1990; Kim et al., 1996; Lenz et al., 1992), esters (Scholz et al., 1994), epoxides (Bear et al., 1997), aromatic rings (Aróstegui et al., 1999; Kim et al., 1991; Takagi et al.,

1999), and nitrile groups (Lenz et al., 1992). Although these are fascinating examples of the diversity of PHAs, several more recent advances in functionalization have great potential for commercial significance and more advances applications.

1.5.1 Fluorination of the PHA α-carbon

Fluoropolymers are incredibly useful materials that have transformed the fields of polymer chemistry and materials science, with the classic example being that of poly(tetrafluoroethylene)

(PTFE) which has extraordinary physical properties (Plunkett, 1941). These properties arise from the C-F bond, which is the strongest single bond to carbon that can be made, and the unique electronic interactions that result from the high electronegativity of fluorine (O’Hagan, 2008).

Organofluorine compounds have also been used extensively in the pharmaceutical industry, where substitutions with fluorine often dramatically alter enzymatic interactions (Kirsch, 2005; Purser et al., 2008). This utility is enhanced by the dearth of organofluorine compounds found in nature; very few metabolic pathways exist that can manipulate C-F bonds, and there are a limited number

29 of biologically produced organofluorine structures (Dong et al., 2004; Leong et al., 2017; Twigg and Socha, 2001; Walker and Bong Chui Lien, 1981).

Recently, a biosynthetic approach for producing fluorinated PHAs was reported by

Thuronyi et al. (2017). This novel recombinant pathway has been included here as an excellent example of the chemical modification of existing PHAs through non-synthetic means (Fig. 1.12), and because there is increasing interest in developing chemistries for modification of the notoriously unreactive C-F bond.

30

Figure 1.12: Recombinant pathway for 2-fluoro-PHA synthesis in Escherichia coli. Using biologically available fluoromalonate as a feedstock, poly[2-fluoro-(R)-3-hydroxybutyrate] can be synthesized as a copolymer with (R)-3-hydroxybutyrate comonomers (Thuronyi et al., 2017). 31

Exogenous fluoromalonate is imported into the cell by one of several potential transporter systems; Thuronyi et al. found that the most effective system was the two-component

Na+:malonate symporter MadLM isolated from Pseudomonas fluorescens (Schaffitzel et al., 1998;

Thuronyi et al., 2017). The intracellular fluoromalonate is then converted to fluoromalonyl coenzyme A (fluoromalonyl-CoA) in a two-step reaction by the malonate-CoA ligase MatB isolated from Rhodopseudomonas palustris (Crosby et al., 2012). The unique acetoacetyl-CoA synthase NphT7 isolated from Streptomyces sp. CL190 catalyzes the condensation of one molecule of acetyl-CoA with one molecule of fluoromalonyl-CoA to yield one molecule of acetofluoroacetyl-CoA concomitant with the release of CO2 and free coenzyme A (Okamura et al.,

2010; Walker et al., 2013). PhaB, an acetoacetyl-CoA reductase isolated from Ralstonia eutropha

(Peoples and Sinskey, 1989a), catalyzes the chemoselective reduction of the beta keto group of acetofluoroacetyl-CoA to yield 2-fluoro-(R)-3-hydroxybutyryl-CoA. This substrate, along with the non-fluorinated analog (R)-3-hydroxybutyrate which occurs due to the presence of the key fatty acid biosynthesis metabolite malonyl-CoA, is finally polymerized by the PHA synthase PhaC isolated from Ralstonia eutropha (Peoples and Sinskey, 1989b) to generate the copolymer poly[2- fluoro-(R)-3-hydroxybutyrate-co-(R)-3-hydroxybutyrate]. This is the first described biosynthesis of a fluoro-PHA, and represents a novel approach to diversifying the portfolio of chemically modified PHAs by the manipulation of the ɑ-carbon.

With the increasing number of synthetic organofluorine compounds, driven in large part by the development of pharmaceuticals, there has been increasing interest in the activation and scission of C-F bonds for further modification. This is a challenge however, given the strength of the C-F bond and its unsuitability as a leaving group (O’Hagan, 2008), which necessitates the use

32 of harsh reaction conditions such as strong nucleophiles (e.g. organolithium reagents) and high temperatures (Fuchibe et al., 2009; Perutz and Braun, 2007; Pigeon et al., 2010). More recently, mild reaction conditions have been found that exploit the fluorophilicity of silicon to cleave non- activated, aliphatic C-F bonds (Tanaka et al., 2016).

Defluorosilylation of alkyl fluorides with relatively mild conditions has been recently achieved independently by several groups (Cui et al., 2018; Liu et al., 2019; Mallick et al., 2019).

Two examples of these defluorosilylation reactions with alkyl fluorides are shown in Fig. 1.13. It is possible that these reaction conditions could be applied to poly[2-fluoro-(R)-3-hydroxybutyrate] in which the fluorine is in an activated position on the α-carbon, allowing further modification of the organosilicon moiety (Komiyama et al., 2017). However, as the ester linkage in PHAs exhibits poor acid/base resistance, more research is needed to determine whether these new methods employed for defluorosilylation are capable of cleaving the C-F bond without destruction of the polymer. Chemical modification of the PHA α-carbon, if possible, would offer an interesting

33 alternative to the current portfolio of functionalizations, most of which utilize reactive groups on the branch chain.

Figure 1.13: Recent methods for base-mediated defluorosilylation. (A) Two examples of independently-devised methods for defluorosilylation of alkyl fluorides (Cui et al., 2018; Liu et al., 2019). (B) It may be possible to adapt these methods for the modification of 2-fluoro-PHA.

An important aspect to consider in the development of fluoropolymers is their environmental impact. Perfluoropolymers such as poly(tetrafluoroethylene) are persistent in the environment, owing to their high stability and chemical inertness. However, many of these materials have been found to be nontoxic and meet the criteria for “polymers of low concern,” which has been extensively reviewed by Henry et al. (2018). The primary concern of these materials, with respect to environmental toxicology, is the toxicity of perfluoroalkyl acids

(PFAAs) which are used as surfactants for fluoropolymer synthesis, or in the case of side-chain fluorinated polymers can be generated by polymer degradation (Liu and Mejia Avendaño, 2013).

PFAAs are regarded as persistent bioaccumulative toxins and have been shown to be carcinogenic

(Takacs and Abbott, 2007), hepatotoxic (Seacat et al., 2002), and teratogenic (Wolf et al., 2007)

34 in a variety of mammalian studies. These compounds have also been implicated in thyroid disease in humans (Melzer David et al., 2010; Winquist and Steenland, 2014), and these combined toxicological effects make it imperative to prevent environmental pollution with PFAAs and potential fluoropolymer precursors. In the case of α-fluorinated PHAs, there are documented pathways of microbial defluorination that may be capable of maintaining the biodegradability typical of PHA polymers, however the toxic potential of 2-fluorobutyrate and its degradation products must still be investigated (Leong et al., 2017; Walker and Bong Chui Lien, 1981). For these reasons, we should exercise caution in the development of fluoro-PHAs.

1.5.2 Development and biosynthesis of azido-PHAs

Click reactions comprise a group of synthetic, chemoselective strategies that have no side products, occur under mild reaction conditions, and proceed rapidly with high yields (Kolb et al.,

2001). Considered the paradigm among click reactions is the so-called copper alkyne-azide cycloaddition reaction (CuAAC), which reliably yields 1,2,3-triazole moieties under mild conditions (Fig. 1.14) (Rostovtsev et al., 2002). The reaction involves a thermodynamically- favored non-concerted cycloaddition featuring Cu(I) as the reactive coordination center by which alkynes and organoazides rearrange to regioselectively form triazole heterocycles. The extraordinary chemoselectivity of the CuAAC and its recognized value as a model reaction in pharmaceutical applications has led to an expanded availability of commercial sources of azido- or alkynyl-conjugated biomolecules. More recently, strain-promoted variants of the alkyne-azide cycloaddition (SPAAC) were developed, with substituted cyclooctynes as their centerpiece, born out of the desire to employ these popular reactions in organisms, avoid the use of deleterious copper, and preserve the reaction’s kinetic parameters in vivo (Agard et al., 2004; Baskin et al.,

2007). Progress in the budding field of orthogonal and bioorthogonal chemical transformations

35 has been rapidly developing and has produced examples which include, but are not limited to; UV- promoted thiol-ene click reactions between thiols and alkenes/alkynes (Dondoni, 2008; Hoyle and

Bowman, 2010), (4+2) cycloadditions between dienes and dienophiles such as normal and reverse- electron-demand Diels-Alder reactions (Pipkorn et al., 2009; Shi and Shoichet, 2008), and the

Staudinger ligation between organoazides and organophosphorus reagents (Saxon and Bertozzi,

2000; Schilling et al., 2011).

36

Figure 1.14: Examples of how click reactions can be used to expand functionality of PHA polymers. In addition to the copper catalyzed azide-alkyne cycloaddition reaction, thiol-ene click reactions and copper-free cycloadditions with strained alkynes are also possible.

In view of the need to develop simple and efficient solutions to the chemical modification of PHAs, Nomura et al. led an effort to produce PHAMCL polymers containing sidechains with reactive organoazide functional groups (Pinto et al., 2016). This modification would allow researchers to bridge the latent reactivity of these polymers with a vast library of click-chemistry

37 based reagents, commercial and otherwise, and offers the potential to expand the reach of PHA applications into high-value areas such as biomedicine, advanced materials, and others.

The implementation of a bacterial strain like E. coli LSBJ, able to uptake and redirect fatty acid feedstocks towards the production of PHA as described previously in this chapter, inspired the development of a short synthetic route to produce chemically tractable azidofatty acids. For this purpose, 8 and 12-carbon α,ω-diols were converted in three steps to the corresponding ω- azidofatty acids in approximately 35% overall yields (Pinto et al., 2016). In addition to the ω- azidofatty acids, 10-azidodecanoic acid was synthesized from commercially available 10- bromodecanoic acid in a single step and with 87% yield. The metabolic incorporation of the synthetic azidofatty acids into PHA copolymers was assessed by fermentation studies employing

E. coli LSBJ and octanoic acid as a primary monomer. E. coli LSBJ was able to uptake all three azidofatty acids, leading to the production of copolymers of variable % azide compositions. Of the materials obtained, poly-[(R)-3-hydroxyoctanoate-co-10-azido-(R)-3-hydroxydecanoate] was employed by the researchers in SPAAC conjugation experiments with the commercially-available cyclooctyne (1R,8S,9s)-bicyclo[6.1.0]non-4-yn-9-ylmethanol.

Nomura et al.’s exploratory study provided a clear answer about the feasibility of introducing organoazides into PHA copolymers and using strain-promoted click chemistry conditions to modify them. More recently, Scholz et al. expanded on this work by providing examples of how copper catalyzed click chemistry conditions could be applied to the modification of PHAs (Nkrumah-Agyeefi and Scholz, 2017). Their study centered on the evaluation of two distinct yet functionally complementary copolymers with nonanoate as the primary monomer: one displaying an azide, and another an alkyne functional group (Fig. 1.15). Optimized production of the alkyne-containing polyester was performed by fermentation with P. oleovorans ATCC 29347

38 in the presence of a 1:1 molar ratio of sodium nonanoate and 10-undecynoic acid (a 40 mM total concentration), and 10 mM sodium acetate over 20h. As opposed to employing synthetic azidofatty acid feedstocks to fermentatively incorporate into PHA, Scholz et al reported an alternative approach that involved the production of a brominated PHA intermediate to be subsequently converted to AzidoPHA by nucleophilic substitution with NaN3. Incorporation of brominated fatty acids into PHA was confirmed by 1H NMR as represented by the methylene signal at 3.41 ppm (-

CH2Br). The brominated side chains were converted by substitution to organoazides by reaction with sodium azide in DMF and their presence confirmed by their characteristic FTIR stretch at

2093 cm-1.

Figure 1.15: Examples of alkynyl-PHA and bromo-PHA biosynthesis by Pseudomonas oleovorans. The azido-PHA copolymers obtained were modified by copper-catalyzed click chemistry.

The authors emphasized the need to test methodologies for this step as they found difficulties when attempting to use a single copper-catalyzed method across polymer systems. The alkynyl-PHA copolymer was conjugated with methyl-2-azidoacetate in the presence of CuBr(PPh3)3, a copper catalyst soluble in tetrahydrofuran (THF) (Malkoch et al., 2005). The reaction progress was monitored by the disappearance of the stretch at 3292 cm-1, representative of terminal alkynes in

39

FTIR spectroscopy. Formation of the triazole adduct was confirmed by 1H NMR spectroscopy. An analogous method was employed in the click chemistry reaction of azido-PHA copolymer with propargyl benzoate. The key difference was the use of the more common click chemistry conditions of CuSO4 5H2O as catalyst in an aqueous/THF solvent mixture containing sodium ascorbate. Under similar reaction conditions the conjugation of azido-PHA with propargyl acetate failed to yield product.

1.5.3 Biosynthesis and modifications of unsaturated PHAs

One type of chemically modifiable monomer unit that has been incorporated into PHAs are ω-unsaturated fatty-acids of varying chain lengths (e.g. 10-undecenoic acid and 7-octenoic acid). The terminal alkenes featured in these polymers have become an attractive chemical handle for their modification post polymerization, mainly by oxidative cleavage reactions. These strategies have been the subject of much scrutiny and have been featured and well described elsewhere (Kai and Loh, 2014).

This notwithstanding, a few salient examples have surfaced in recent years such as that by

Scholz et al. which showed that PHAs containing 10-undecenoate repeating units can be modified to produce ionizable and hydrophilic PHAs (Sparks and Scholz, 2008). In their study, the terminal alkene moieties in poly-[(R)-3-hydroxyoctanoate-co-(R)-3-hydroxyundecenoate] were epoxidized by treatment with 3-chloroperbenzoic acid and ring-opened by nucleophilic reaction with diethanolamine. The resulting ionizable PHA polymer was characterized by 1D and 2D NMR spectroscopy and solubility tests confirmed its miscibility in water. The polycationic character of the modified PHA polymer served as the basis for investigating its value as a plasmid DNA delivery system. The authors demonstrated the polymer’s ability to tightly bind and protect plasmid DNA from endonuclease degradation (Sparks and Scholz, 2009).

40

Other novel approaches for the modification and application of unsaturated PHA polymers have been performed by employing thiol-ene click chemistry. This type of click reaction involves the radical-driven covalent coupling of alkene and thiol functional groups in the presence of ultraviolet light (UV) and in the presence of a photoinitiator (Kade et al., 2010). This reaction has been successfully demonstrated by Levine et al. using poly(3-hydroxybutyrate-co-3-hydroxy-10- undecenoate) (PHBUe) copolymers biosynthesized by E. coli LSBJ (2015). Copolymers were cross-linked using pentaerythritol tetrakis(3-mercaptopropionate) (PETMP) and the radical photoinitiator 2,2-dimethoxy-2-phenylacetophenone (DMPA) (Fig. 1.16), resulting in a 200% increase in tensile strength. Cytotoxicity assays conducted with human mesenchymal stem cells

(hMSC) showed that cell viability was unaffected by the cross-linking, with hMSC-seeded PHBUe and PHBUX scaffolds exhibiting no difference between each other or a positive control. This represents an exciting new PHA material with potential use in tissue engineering.

Figure 1.16: Cross-linked poly(3-hydroxybutyrate-co-3-hydroxy-10-undecenoate) (PHBUe) produced via thiol-ene click chemistry.

The potential use of PHAs in biomedical applications such as drug delivery and tissue engineering has generally been limited by the methodologies available to chemically modify and

41 fine-tune their properties (Hazer et al., 2012; Poletto et al., 2008). Recent advances in the biosynthesis of azide-, fluorine-, and alkyne-containing PHAs are an excellent first step towards the production of useful chemically modifiable materials.

1.6 Objectives of this dissertation

A number of PHAs have been studied in great detail as biodegradable alternatives to conventional plastics, particularly SCL PHAs such as PHB and PHBV or the SCL-co-MCL polymer PHBHHx, which can be more readily produced at an industrial scale (Chanprateep, 2010;

Noda et al., 2010). MCL PHAs and diverse SCL-co-MCL PHAs offer exciting new material properties and chemical handles for functionalization; however, controlled biosynthesis of these polymers has only recently been achieved and is hindered by low yields and prohibitive production costs (Chen and Albertsson, 2019; Tappel et al., 2012b).

MCL PHA biosynthesis in recombinant E. coli LSBJ most commonly occurs via β- oxidation intermediates, and can produce a variety of MCL homopolymers and copolymers with controlled compositions. This process is limited in its production capacity due to the poor growth of E. coli on fatty acid substrates and the limiting effect of catabolite repression when supplemented with a preferred carbon source such as glucose (Deutscher, 2008; Feng and Cronan,

2012; Nunn, 1986). The first objective of this dissertation is to evaluate the role of transcriptional regulators of fatty acid catabolism in the PHA biosynthesis pathway of E. coli LSBJ, with a focus on enhancing MCL PHA yields (Chapter 2). The hypothesis presented in this chapter is that the deletion of the negative transcriptional regulators arcA and ompR from E. coli LSBJ will increase expression of the fad genes necessary for PHA biosynthesis from fatty and improve polymer yields.

42

In a separate effort to address the prohibitive cost of PHA biosynthesis, the second objective of this dissertation is to explore alternative lignocellulosic feedstocks as cheaper alternatives to pure sugars (Chapter 3). Using a combination of the de novo biosynthesis pathway

(Fig. 1.4) from acetyl-CoA and PhaG-mediated MCL monomer incorporation, both PHB and

PHB-co-MCL PHAs were successfully produced from hydrolyzed paper mill waste pulp.

Biosynthesis of many desired chemicals and materials can be achieved in high yields using fed-batch, high-density fermentation processes in a scalable vessel such as a stirred tank reactor

(Lau et al., 2004; Lee, 1996; Riesenberg and Guthke, 1999). These techniques have been applied to PHA biosynthesis as well, resulting in significant improvements to productivity primarily for

SCL polymers (Amini et al., 2020; Blunt et al., 2018; Koller, 2018; Lee et al., 1999). However, there have been comparatively few studies to investigate the production of MCL PHAs using high- density fermentation, and fewer still that have maintained control over monomer composition

(Follonier et al., 2015; Gao et al., 2018; Gopi et al., 2018; Jiang et al., 2013). The work presented in Chapter 4 describes a process optimization for the biosynthesis of MCL PHAs with controlled composition via high-density fed-batch fermentation with E. coli LSBJ. The objective of this work was to significantly improve MCL PHA yields from those observed in previous studies that relied on unscalable shake flask culture techniques, including the biosynthesis of functional azido PHA polymers.

Chapter 5 expands on the novel azido PHAs produced in Chapter 4 and in previous work by Pinto et al. to develop functionalized polymeric nanoparticles for targeted drug delivery (2016).

This chapter presents promising preliminary data for the production of endotoxin-free nanoparticles capable of encapsulating the chemotherapeutic drug paclitaxel at a high efficiency.

43

1.7 References

Agard, N.J., Prescher, J.A., Bertozzi, C.R., 2004. A Strain-Promoted [3 + 2] Azide−Alkyne

Cycloaddition for Covalent Modification of Biomolecules in Living Systems. J. Am.

Chem. Soc. 126, 15046–15047. https://doi.org/10.1021/ja044996f

Aiba, H., Mizuno, T., 1990. Phosphorylation of a bacterial activator protein, OmpR, by a protein

kinase, EnvZ, stimulates the transcription of the ompF and ompC genes in Escherichia

coli. FEBS Letters 261, 19–22. https://doi.org/10.1016/0014-5793(90)80626-T

Amini, M., Yousefi-Massumabad, H., Younesi, H., Abyar, H., Bahramifar, N., 2020. Production

of the polyhydroxyalkanoate biopolymer by Cupriavidus necator using beer brewery

wastewater containing maltose as a primary carbon source. Journal of Environmental

Chemical Engineering 8, 103588. https://doi.org/10.1016/j.jece.2019.103588

Aróstegui, S.M., Aponte, M.A., Díaz, E., Schröder, E., 1999. Bacterial Polyesters Produced by

Pseudomonas oleovorans Containing Nitrophenyl Groups. Macromolecules 32, 2889–

2895. https://doi.org/10.1021/ma981482q

Asrar, J., Valentin, H.E., Berger, P.A., Tran, M., Padgette, S.R., Garbow, J.R., 2002. Biosynthesis

and properties of poly(3-hydroxybutyrate-co-3-hydroxyhexanoate) polymers.

Biomacromolecules 3, 1006–1012. https://doi.org/10.1021/bm025543a

Barboza, L.G.A., Dick Vethaak, A., Lavorante, B.R.B.O., Lundebye, A.-K., Guilhermino, L.,

2018. Marine microplastic debris: An emerging issue for food security, food safety and

human health. Marine Pollution Bulletin 133, 336–348.

https://doi.org/10.1016/j.marpolbul.2018.05.047

44

Barham, P.J., Barker, P., Organ, S.J., 1992. Physical properties of poly(hydroxybutyrate) and

copolymers of hydroxybutyrate and hydroxyvalerate. FEMS Microbiology Letters 103,

289–298. https://doi.org/10.1016/0378-1097(92)90322-F

Baskin, J.M., Prescher, J.A., Laughlin, S.T., Agard, N.J., Chang, P.V., Miller, I.A., Lo, A., Codelli,

J.A., Bertozzi, C.R., 2007. Copper-free click chemistry for dynamic in vivo imaging. Proc.

Natl. Acad. Sci. U.S.A. 104, 16793–16797. https://doi.org/10.1073/pnas.0707090104

Bear, M.-M., Leboucher-Durand, M.-A., Langlois, V., Lenz, R.W., Goodwin, S., Guérin, P., 1997.

Bacterial poly-3-hydroxyalkenoates with epoxy groups in the side chains. Reactive and

Functional Polymers, International Conference of Applied Physical Chemistry 34, 65–77.

https://doi.org/10.1016/S1381-5148(97)00024-2

Beutin, L., Achtman, M., 1979. Two Escherichia coli chromosomal cistrons, sfrA and sfrB, which

are needed for expression of F factor tra functions. J. Bacteriol. 139, 730–737.

Beutin, L., Manning, P.A., Achtman, M., Willetts, N., 1981. sfrA and sfrB products of Escherichia

coli K-12 are transcriptional control factors. J. Bacteriol. 145, 840–844.

Black, P.N., 1990. Characterization of FadL-specific fatty acid binding in Escherichia coli.

Biochimica et Biophysica Acta (BBA) - Lipids and Lipid Metabolism 1046, 97–105.

https://doi.org/10.1016/0005-2760(90)90099-J

Black, P.N., DiRusso, C.C., 1994. Molecular and biochemical analyses of fatty acid transport,

metabolism, and gene regulation in Escherichia coli. Biochim. Biophys. Acta 1210, 123–

145. https://doi.org/10.1016/0005-2760(94)90113-9

Blunt, W., Levin, D.B., Cicek, N., 2018. Bioreactor Operating Strategies for Improved

Polyhydroxyalkanoate (PHA) Productivity. Polymers 10, 1197.

https://doi.org/10.3390/polym10111197

45

Brennan, R.G., 1993. The winged-helix DNA-binding motif: Another helix-turn-helix takeoff.

Cell 74, 773–776. https://doi.org/10.1016/0092-8674(93)90456-Z

Buxton, R.S., Drury, L.S., 1984. Identification of the dye gene product, mutational loss of which

alters envelope protein composition and also affects sex factor F expression in Escherichia

coli K-12. Molec. Gen. Genet. 194, 241–247. https://doi.org/10.1007/BF00383523

Buxton, R.S., Drury, L.S., 1983. Cloning and insertional inactivation of the dye (sfrA) gene,

mutation of which affects sex factor F expression and dye sensitivity of Escherichia coli

K-12. Journal of bacteriology 154, 1309–1314.

Buxton, R.S., Drury, L.S., Curtis, C.A., 1983. Dye sensitivity correlated with envelope protein

changes in dye (sfrA) mutants of Escherichia coli K12 defective in the expression of the

sex factor F. Journal of general microbiology 129, 3363–3370.

Buxton, R.S., Hammer-Jespersen, K., Hansen, T.D., 1978. Insertion of bacteriophage lambda into

the deo operon of Escherichia coli K-12 and isolation of plaque-forming lambdadeo+

transducing bacteriophages. J. Bacteriol. 136, 668–681.

Campbell, J.W., Cronan, J.E., 2002. The enigmatic Escherichia coli fadE gene is yafH. J.

Bacteriol. 184, 3759–3764.

Campbell, J.W., Morgan-Kiss, R.M., Cronan, J.E., 2003. A new Escherichia coli metabolic

competency: growth on fatty acids by a novel anaerobic beta-oxidation pathway. Mol.

Microbiol. 47, 793–805. https://doi.org/10.1046/j.1365-2958.2003.03341.x

Chanprateep, S., 2010. Current trends in biodegradable polyhydroxyalkanoates. J. Biosci. Bioeng.

110, 621–632. https://doi.org/10.1016/j.jbiosc.2010.07.014

Chen, G.-Q., Albertsson, A.-C., 2019. Polyhydroxyalkanoates and Other Biopolymers.

Biomacromolecules 20, 3211–3212. https://doi.org/10.1021/acs.biomac.9b01044

46

Cho, B.K., Knight, E.M., Palsson, B.O., 2006. Transcriptional regulation of the fad regulon genes

of Escherichia coli by ArcA. Microbiology 152, 2207–2219. https://doi.org/152/8/2207

[pii]

Cronan, J.E.Jr., Subrahmanyam, S., 1998. FadR, transcriptional co-ordination of metabolic

expediency. Molecular Microbiology 29, 937–943. https://doi.org/10.1046/j.1365-

2958.1998.00917.x

Crosby, H.A., Rank, K.C., Rayment, I., Escalante-Semerena, J.C., 2012. Structure-guided

expansion of the substrate range of methylmalonyl coenzyme A synthetase (MatB) of

Rhodopseudomonas palustris. Appl Environ Microbiol 78, 6619–6629.

https://doi.org/10.1128/AEM.01733-12

Cui, B., Jia, S., Tokunaga, E., Shibata, N., 2018. Defluorosilylation of fluoroarenes and

fluoroalkanes. Nat Commun 9, 1–8. https://doi.org/10.1038/s41467-018-06830-w

Deutscher, J., 2008. The mechanisms of carbon catabolite repression in bacteria. Curr. Opin.

Microbiol. 11, 87–93. https://doi.org/10.1016/j.mib.2008.02.007

DiRusso, C.C., Tsvetnitsky, V., Højrup, P., Knudsen, J., 1998. Fatty acyl-CoA binding domain of

the transcription factor FadR. Characterization by deletion, affinity labeling, and

isothermal titration calorimetry. J. Biol. Chem. 273, 33652–33659.

https://doi.org/10.1074/jbc.273.50.33652

Doi, Y., Abe, C., 1990. Biosynthesis and characterization of a new bacterial copolyester of 3-

hydroxyalkanoates and 3-hydroxy-ω-chloroalkanoates. Macromolecules 23, 3705–3707.

https://doi.org/10.1021/ma00217a027

47

Dondoni, A., 2008. The emergence of thiol-ene coupling as a click process for materials and

bioorganic chemistry. Angew. Chem. Int. Ed. Engl. 47, 8995–8997.

https://doi.org/10.1002/anie.200802516

Dong, C., Huang, F., Deng, H., Schaffrath, C., Spencer, J.B., O’Hagan, D., Naismith, J.H., 2004.

Crystal structure and mechanism of a bacterial fluorinating enzyme. Nature 427, 561–565.

https://doi.org/10.1038/nature02280

Dong, H., Liffland, S., Hillmyer, M.A., Chang, M.C.Y., 2019. Engineering in Vivo Production of

α-Branched Polyesters. J. Am. Chem. Soc. 141, 16877–16883.

https://doi.org/10.1021/jacs.9b08585

Dong, T., Yu, R., Schellhorn, H., 2011. Antagonistic regulation of motility and transcriptome

expression by RpoN and RpoS in Escherichia coli. Mol. Microbiol. 79, 375–386.

https://doi.org/10.1111/j.1365-2958.2010.07449.x

Drury, L.S., Buxton, R.S., 1985. DNA sequence analysis of the dye gene of Escherichia coli

reveals amino acid homology between the dye and OmpR proteins. J. Biol. Chem. 260,

4236–4242.

Eriksen, M., Lebreton, L.C.M., Carson, H.S., Thiel, M., Moore, C.J., Borerro, J.C., Galgani, F.,

Ryan, P.G., Reisser, J., 2014. Plastic Pollution in the World’s Oceans: More than 5 Trillion

Plastic Pieces Weighing over 250,000 Tons Afloat at Sea. PLoS ONE 9, e111913.

https://doi.org/10.1371/journal.pone.0111913

Feng, Y., Cronan, J.E., 2012. Crosstalk of Escherichia coli FadR with Global Regulators in

Expression of Fatty Acid Transport Genes. PLoS ONE 7, e46275.

https://doi.org/10.1371/journal.pone.0046275

48

Fic, E., Bonarek, P., Gorecki, A., Kedracka-Krok, S., Mikolajczak, J., Polit, A., Tworzydlo, M.,

Dziedzicka-Wasylewska, M., Wasylewski, Z., 2009. cAMP Receptor Protein from

Escherichia coli as a Model of Signal Transduction in Proteins - A Review. Journal of

Molecular Microbiology and Biotechnology 17, 1–11. https://doi.org/10.1159/000178014

Finn, R.D., Bateman, A., Clements, J., Coggill, P., Eberhardt, R.Y., Eddy, S.R., Heger, A.,

Hetherington, K., Holm, L., Mistry, J., Sonnhammer, E.L.L., Tate, J., Punta, M., 2014.

Pfam: the protein families database. Nucleic Acids Res. 42, D222-230.

https://doi.org/10.1093/nar/gkt1223

Follonier, S., Riesen, R., Zinn, M., 2015. Pilot-scale Production of Functionalized mcl-PHA from

Grape Pomace Supplemented with Fatty Acids. Chemical and Biochemical Engineering

Quarterly 29, 113–121. https://doi.org/10.15255/CABEQ.2014.2251

Foo, Y.H., Gao, Y., Zhang, H., Kenney, L.J., 2015. Cytoplasmic sensing by the inner membrane

histidine kinase EnvZ. Progress in Biophysics and Molecular Biology, Mechanism of

Transmembrane Signaling 118, 119–129.

https://doi.org/10.1016/j.pbiomolbio.2015.04.005

Forst, S.A., Roberts, D.L., 1994. Signal transduction by the EnvZ-OmpR phosphotransfer system

in bacteria. Res. Microbiol. 145, 363–373.

Franchini, A.G., Ihssen, J., Egli, T., 2015. Effect of Global Regulators RpoS and Cyclic-AMP/CRP

on the Catabolome and Transcriptome of Escherichia coli K12 during Carbon- and Energy-

Limited Growth. PLoS ONE 10, e0133793. https://doi.org/10.1371/journal.pone.0133793

Fuchibe, K., Kaneko, T., Mori, K., Akiyama, T., 2009. Expedient Synthesis of N-Fused Indoles:

A C-F Activation and C-H Insertion Approach. Angewandte Chemie International Edition

48, 8070–8073. https://doi.org/10.1002/anie.200901986

49

Gagnon, K.D., Lenz, R.W., Farris, R.J., Fuller, R.C., 1992. Crystallization behavior and its

influence on the mechanical properties of a thermoplastic elastomer produced by

Pseudomonas oleovorans. Macromolecules 25, 3723–3728.

https://doi.org/10.1021/ma00040a018

Gao, J., Vo, M.T., Ramsay, J.A., Ramsay, B.A., 2018. Overproduction of MCL-PHA with high 3-

hydroxydecanoate Content. Biotechnology and Bioengineering 115, 390–400.

https://doi.org/10.1002/bit.26474

Georgellis, D., Kwon, O., De Wulf, P., Lin, E.C., 1998. Signal decay through a reverse

phosphorelay in the Arc two-component signal transduction system. J. Biol. Chem. 273,

32864–32869.

Georgellis, D., Kwon, O., Lin, E.C., 2001. Quinones as the redox signal for the arc two-component

system of bacteria. Science 292, 2314–2316.

Gogotov, I.N., Gerasin, V.A., Knyazev, Ya.V., Antipov, E.M., Barazov, S.Kh., 2010. Composite

biodegradable materials based on polyhydroxyalkanoate. Appl Biochem Microbiol 46,

607–613. https://doi.org/10.1134/S0003683810060086

Gopi, S., Kontopoulou, M., Ramsay, B.A., Ramsay, J.A., 2018. Manipulating the structure of

medium-chain-length polyhydroxyalkanoate (MCL-PHA) to enhance thermal properties

and crystallization kinetics. Int. J. Biol. Macromol. 119, 1248–1255.

https://doi.org/10.1016/j.ijbiomac.2018.08.016

Hassan, M.K., Abou-Hussein, R., Zhang, X., Mark, J.E., Noda, I., 2006. Biodegradable

Copolymers of 3-Hydroxybutyrate-co-3-Hydroxyhexanoate (NodaxTM), Including Recent

Improvements in their Mechanical Properties. Molecular Crystals and Liquid Crystals 447,

23/[341]-44/[362]. https://doi.org/10.1080/15421400500380028

50

Hathaway, B.G., Bergquist, P.L., 1973. Temperature-sensitive mutations affecting the replication

of F-prime factors in Escherichia coli K12. Molec. Gen. Genet. 127, 297–306.

https://doi.org/10.1007/BF00267100

Hazer, D.B., Kılıçay, E., Hazer, B., 2012. Poly(3-hydroxyalkanoate)s: Diversification and

biomedical applications. Materials Science and Engineering: C 32, 637–647.

https://doi.org/10.1016/j.msec.2012.01.021

Hein, S., Söhling, B., Gottschalk, G., Steinbüchel, A., 1997. Biosynthesis of poly(4-

hydroxybutyric acid) by recombinant strains of Escherichia coli. FEMS Microbiol. Lett.

153, 411–418. https://doi.org/10.1111/j.1574-6968.1997.tb12604.x

Henry, B.J., Carlin, J.P., Hammerschmidt, J.A., Buck, R.C., Buxton, L.W., Fiedler, H., Seed, J.,

Hernandez, O., 2018. A critical review of the application of polymer of low concern and

regulatory criteria to fluoropolymers. Integrated Environmental Assessment and

Management 14, 316–334. https://doi.org/10.1002/ieam.4035

Henry, M.F., Cronan, J.E., 1992. A new mechanism of transcriptional regulation: Release of an

activator triggered by small molecule binding. Cell 70, 671–679.

https://doi.org/10.1016/0092-8674(92)90435-F

Higashitani, A., Nishimura, Y., Hara, H., Aiba, H., Mizuno, T., Horiuchi, K., 1993.

Osmoregulation of the fatty acid receptor gene fadL in Escherichia coli. Mol.Gen.Genet.

240, 339–347.

Hoyle, C.E., Bowman, C.N., 2010. Thiol–Ene Click Chemistry. Angewandte Chemie International

Edition 49, 1540–1573. https://doi.org/10.1002/anie.200903924

51

Iram, S.H., Cronan, J.E., 2006. The β-Oxidation Systems of Escherichia coli and Salmonella

enterica Are Not Functionally Equivalent. Journal of Bacteriology 188, 599–608.

https://doi.org/10.1128/JB.188.2.599-608.2006

Iram, S.H., Cronan, J.E., 2005. Unexpected Functional Diversity among FadR Fatty Acid

Transcriptional Regulatory Proteins. J. Biol. Chem. 280, 32148–32156.

https://doi.org/10.1074/jbc.M504054200

Ishige, K., Nagasawa, S., Tokishita, S.-I., Mizuno, T., 1994. A novel device of bacterial signal

transducers. The EMBO journal 13, 5195.

Itou, H., Tanaka, I., 2001. The OmpR-Family of Proteins: Insight into the Tertiary Structure and

Functions of Two-Component Regulator Proteins. The Journal of Biochemistry 129, 343–

350.

Iuchi, S., Lin, E.C., 1992. Purification and phosphorylation of the Arc regulatory components of

Escherichia coli. Journal of bacteriology 174, 5617–5623.

Iuchi, S., Lin, Ec., 1988. arcA (dye), a global regulatory gene in Escherichia coli mediating

repression of enzymes in aerobic pathways. Proceedings of the National Academy of

Sciences 85, 1888–1892.

Jamieson, A.F., Bergquist, P.L., 1976. Genetic mapping of chromosomal mutations affecting the

replication of the F-factor of Escherichia coli. Molec. Gen. Genet. 148, 221–223.

https://doi.org/10.1007/BF00268388

Jiang, F., An, C., Bao, Y., Zhao, X., Jernigan, R.L., Lithio, A., Nettleton, D., Li, L., Wurtele, E.S.,

Nolan, L.K., Lu, C., Li, G., 2015. ArcA Controls Metabolism, Chemotaxis, and Motility

Contributing to the Pathogenicity of Avian Pathogenic Escherichia coli. Infection and

Immunity 83, 3545–3554. https://doi.org/10.1128/IAI.00312-15

52

Jiang, X.J., Sun, Z., Ramsay, J.A., Ramsay, B.A., 2013. Fed-batch production of MCL-PHA with

elevated 3-hydroxynonanoate content. AMB Express 3, 50. https://doi.org/10.1186/2191-

0855-3-50

Kade, M.J., Burke, D.J., Hawker, C.J., 2010. The power of thiol-ene chemistry. Journal of Polymer

Science Part A: Polymer Chemistry 48, 743–750. https://doi.org/10.1002/pola.23824

Kai, D., Loh, X.J., 2014. Polyhydroxyalkanoates: Chemical Modifications Toward Biomedical

Applications. ACS Sustainable Chem. Eng. 2, 106–119.

https://doi.org/10.1021/sc400340p

Kameda, K., Nunn, W.D., 1981. Purification and characterization of acyl coenzyme A synthetase

from Escherichia coli. J. Biol. Chem. 256, 5702–5707.

Kato, N., Tsuzuki, M., Aiba, H., Mizuno, T., 1995. Gene activation by the Escherichia coli positive

regulator OmpR: A mutational study of the DNA-binding domain of OmpR. Molec. Gen.

Genet. 248, 399–406. https://doi.org/10.1007/BF02191639

Keseler, I.M., Mackie, A., Peralta-Gil, M., Santos-Zavaleta, A., Gama-Castro, S., Bonavides-

Martínez, C., Fulcher, C., Huerta, A.M., Kothari, A., Krummenacker, M., Latendresse, M.,

Muñiz-Rascado, L., Ong, Q., Paley, S., Schröder, I., Shearer, A.G., Subhraveti, P., Travers,

M., Weerasinghe, D., Weiss, V., Collado-Vides, J., Gunsalus, R.P., Paulsen, I., Karp, P.D.,

2013. EcoCyc: fusing model organism databases with systems biology. Nucl. Acids Res.

41, D605–D612. https://doi.org/10.1093/nar/gks1027

Kim, O., Gross, R.A., Hammar, W.J., Newmark, R.A., 1996. Microbial Synthesis of Poly(β-

hydroxyalkanoates) Containing Fluorinated Side-Chain Substituents. Macromolecules 29,

4572–4581. https://doi.org/10.1021/ma960059j

53

Kim, Y.B., Lenz, R.W., Fuller, R.C., 1991. Preparation and characterization of poly(β-

hydroxyalkanoates) obtained from Pseudomonas oleovorans grown with mixtures of 5-

phenylvaleric acid and n-alkanoic acids. Macromolecules 24, 5256–5260.

https://doi.org/10.1021/ma00019a004

Kirsch, P., 2005. Applications of Organofluorine Compounds, in: Modern Fluoroorganic

Chemistry. John Wiley & Sons, Ltd, pp. 203–277.

https://doi.org/10.1002/352760393X.ch4

Kolb, H.C., Finn, M.G., Sharpless, K.B., 2001. Click Chemistry: Diverse Chemical Function from

a Few Good Reactions. Angewandte Chemie International Edition 40, 2004–2021.

https://doi.org/10.1002/1521-3773(20010601)40:11<2004::AID-ANIE2004>3.0.CO;2-5

Koller, M., 2018. A Review on Established and Emerging Fermentation Schemes for Microbial

Production of Polyhydroxyalkanoate (PHA) Biopolyesters. Fermentation 4, 30.

https://doi.org/10.3390/fermentation4020030

Komiyama, T., Minami, Y., Hiyama, T., 2017. Recent Advances in Transition-Metal-Catalyzed

Synthetic Transformations of Organosilicon Reagents. ACS Catal. 7, 631–651.

https://doi.org/10.1021/acscatal.6b02374

Kundig, W., Ghosh, S., Roseman, S., 1964. Phosphate bound to histidine in a protein as an

intermediate in a novel phospho-transferase system. Proc. Natl. Acad. Sci. U.S.A. 52,

1067–1074. https://doi.org/10.1073/pnas.52.4.1067

Kwon, O., Georgellis, D., Lynch, A.S., Boyd, D., Lin, E.C.C., 2000. The ArcB sensor kinase of

Escherichia coli: genetic exploration of the transmembrane region. Journal of bacteriology

182, 2960–2966.

54

Lageveen, R.G., Huisman, G.W., Preusting, H., Ketelaar, P.E.F., Eggink, G., Witholt, B., 1988.

Formation of Polyesters by Pseudomonas oleovorans: Effect of Substrates on Formation

and Composition of Poly-(R)-3-Hydroxyalkanoates and Poly-(R)-3-Hydroxyalkenoates.

Applied and environmental microbiology. https://doi.org/10.1128/aem.54.12.2924-

2932.1988

Lau, J., Tran, C., Licari, P., Galazzo, J., 2004. Development of a high cell-density fed-batch

bioprocess for the heterologous production of 6-deoxyerythronolide B in Escherichia coli.

J. Biotechnol. 110, 95–103. https://doi.org/10.1016/j.jbiotec.2004.02.001

Lazarevic, D., Aoustin, E., Buclet, N., Brandt, N., 2010. Plastic waste management in the context

of a European recycling society: Comparing results and uncertainties in a life cycle

perspective. Resources, Conservation and Recycling 55, 246–259.

https://doi.org/10.1016/j.resconrec.2010.09.014

Lee, E.Y., Jendrossek, D., Schirmer, A., Choi, C.Y., Steinbüchel, A., 1995. Biosynthesis of

copolyesters consisting of 3-hydroxybutyric acid and medium-chain-length 3-

hydroxyalkanoic acids from 1,3-butanediol or from 3-hydroxybutyrate by Pseudomonas

sp. A33. Appl Microbiol Biotechnol 42, 901–909. https://doi.org/10.1007/BF00191189

Lee, S.Y., 1996. High cell-density culture of Escherichia coli. Trends in Biotechnology 14, 98–

105. https://doi.org/10.1016/0167-7799(96)80930-9

Lee, S.Y., Choi, J., Wong, H.H., 1999. Recent advances in polyhydroxyalkanoate production by

bacterial fermentation: mini-review. International Journal of Biological Macromolecules

25, 31–36. https://doi.org/10.1016/S0141-8130(99)00012-4

Lemoigne, M., 1926. Produits de dehydration et de polymerisation de l’acide ß-oxobutyrique. Bull.

Soc. Chim. Biol. 8, 770–782.

55

Lenz, R.W., Kim, Y.B., Fuller, R.C., 1992. Production of unusual bacterial polyesters by

Pseudomonas oleovorans through cometabolism. FEMS Microbiology Letters 103, 207–

214. https://doi.org/10.1016/0378-1097(92)90311-B

Leong, L.E.X., Khan, S., Davis, C.K., Denman, S.E., McSweeney, C.S., 2017. Fluoroacetate in

plants - a review of its distribution, toxicity to livestock and microbial detoxification. J

Anim Sci Biotechnol 8. https://doi.org/10.1186/s40104-017-0180-6

Lepore, B.W., Indic, M., Pham, H., Hearn, E.M., Patel, D.R., van den Berg, B., 2011. Ligand-

gated diffusion across the bacterial outer membrane. Proc Natl Acad Sci U S A 108, 10121–

10126. https://doi.org/10.1073/pnas.1018532108

Lerner, T.J., Zinder, N.D., 1982. Another gene affecting sexual expression of Escherichia coli.

Journal of bacteriology 150, 156–160.

Levine, A.C., Sparano, A., Twigg, F.F., Numata, K., Nomura, C.T., 2015. Influence of Cross-

Linking on the Physical Properties and Cytotoxicity of Polyhydroxyalkanoate (PHA)

Scaffolds for Tissue Engineering. ACS Biomaterials Science & Engineering 1, 567–576.

https://doi.org/10.1021/acsbiomaterials.5b00052

Li, Z.-J., Qiao, K., Che, X.-M., Stephanopoulos, G., 2017. Metabolic engineering of Escherichia

coli for the synthesis of the quadripolymer poly(glycolate-co-lactate-co-3-

hydroxybutyrate-co-4-hydroxybutyrate) from glucose. Metab. Eng. 44, 38–44.

https://doi.org/10.1016/j.ymben.2017.09.003

Li, Z.-J., Qiao, K., Shi, W., Pereira, B., Zhang, H., Olsen, B.D., Stephanopoulos, G., 2016.

Biosynthesis of poly(glycolate-co-lactate-co-3-hydroxybutyrate) from glucose by

metabolically engineered Escherichia coli. Metabolic Engineering 35, 1–8.

https://doi.org/10.1016/j.ymben.2016.01.004

56

Liu, J., Mejia-Avendaño, S., 2013. Microbial degradation of polyfluoroalkyl chemicals in the

environment: A review. Environment International 61, 98–114.

https://doi.org/10.1016/j.envint.2013.08.022

Liu, X., Wulf, P.D., 2004. Probing the ArcA-P Modulon of Escherichia coli by Whole Genome

Transcriptional Analysis and Sequence Recognition Profiling. J. Biol. Chem. 279, 12588–

12597. https://doi.org/10.1074/jbc.M313454200

Liu, X.W., Zarate, C., Martin, R., 2019. Base-Mediated Defluorosilylation of C(sp2)−F and

C(sp3)−F Bonds. Angewandte Chemie International Edition 58, 2064–2068.

https://doi.org/10.1002/anie.201813294

Lu, J., Tappel, R.C., Nomura, C.T., 2009. Mini-Review: Biosynthesis of Poly(hydroxyalkanoates).

Polymer Reviews 49, 226 248. https://doi.org/10.1080/15583720903048243

Lynch, A.S., Lin, E.C., 1996. Transcriptional control mediated by the ArcA two-component

response regulator protein of Escherichia coli: characterization of DNA binding at target

promoters. Journal of bacteriology 178, 6238–6249.

Maciag, A., Peano, C., Pietrelli, A., Egli, T., De Bellis, G., Landini, P., 2011. In vitro transcription

profiling of the σS subunit of bacterial RNA polymerase: re-definition of the σS regulon

and identification of σS-specific promoter sequence elements. Nucleic Acids Res. 39,

5338–5355. https://doi.org/10.1093/nar/gkr129

Malkoch, M., Schleicher, K., Drockenmuller, E., Hawker, C.J., Russell, T.P., Wu, P., Fokin, V.V.,

2005. Structurally Diverse Dendritic Libraries: A Highly Efficient Functionalization

Approach Using Click Chemistry. Macromolecules 38, 3663–3678.

https://doi.org/10.1021/ma047657f

57

Mallick, S., Xu, P., Würthwein, E.-U., Studer, A., 2019. Silyldefluorination of Fluoroarenes by

Concerted Nucleophilic Aromatic Substitution. Angew. Chem. Int. Ed. Engl. 58, 283–287.

https://doi.org/10.1002/anie.201808646

Malpica, R., Franco, B., Rodriguez, C., Kwon, O., Georgellis, D., 2004. Identification of a

quinone-sensitive redox switch in the ArcB sensor kinase. Proc Natl Acad Sci U S A 101,

13318–13323. https://doi.org/10.1073/pnas.0403064101

Mandin, P., Gottesman, S., 2010. Integrating anaerobic/aerobic sensing and the general stress

response through the ArcZ small RNA. The EMBO journal 29, 3094–3107.

Martin, D.P., Williams, S.F., 2003. Medical applications of poly-4-hydroxybutyrate: a strong

flexible absorbable biomaterial. Biochemical Engineering Journal, Biopolymers 16, 97–

105. https://doi.org/10.1016/S1369-703X(03)00040-8

Matsubara, M., Mizuno, T., 1999. EnvZ-independent phosphotransfer signaling pathway of the

OmpR-mediated osmoregulatory expression of OmpC and OmpF in Escherichia coli.

Biosci. Biotechnol. Biochem. 63, 408–414. https://doi.org/10.1271/bbb.63.408

Matsumoto, K., Kageyama, Y., 2019. Increased Production and Molecular Weight of Artificial

Polyhydroxyalkanoate Poly(2-hydroxybutyrate) Above the Glass Transition Temperature

Threshold. Front. Bioeng. Biotechnol. 7. https://doi.org/10.3389/fbioe.2019.00177

Matsumoto, K., Takase, K., Yamamoto, Y., Doi, Y., Taguchi, S., 2009. Chimeric enzyme

composed of polyhydroxyalkanoate (PHA) synthases from Ralstonia eutropha and

Aeromonas caviae enhances production of PHAs in recombinant Escherichia coli.

Biomacromolecules 10, 682–685. https://doi.org/10.1021/bm801386j

Melzer D., Rice N., Depledge M.H., Henley W.E., Galloway T.S., 2010. Association between

Serum Perfluorooctanoic Acid (PFOA) and Thyroid Disease in the U.S. National Health

58

and Nutrition Examination Survey. Environmental Health Perspectives 118, 686–692.

https://doi.org/10.1289/ehp.0901584

Mizuno, S., Enda, Y., Saika, A., Hiroe, A., Tsuge, T., 2018. Biosynthesis of

polyhydroxyalkanoates containing 2-hydroxy-4-methylvalerate and 2-hydroxy-3-

phenylpropionate units from a related or unrelated carbon source. J. Biosci. Bioeng. 125,

295–300. https://doi.org/10.1016/j.jbiosc.2017.10.010

Mizuno, T., Mizushima, S., 1990. Signal transduction and gene regulation through the

phosphorylation of two regulatory components: the molecular basis for the osmotic

regulation of the porin genes. Molecular Microbiology 4, 1077–1082.

https://doi.org/10.1111/j.1365-2958.1990.tb00681.x

Nikaido, H., 1992. Porins and specific channels of bacterial outer membranes. Mol. Microbiol. 6,

435–442. https://doi.org/10.1111/j.1365-2958.1992.tb01487.x

Nkrumah-Agyeefi, S., Scholz, C., 2017. Chemical modification of functionalized

polyhydroxyalkanoates via “Click” chemistry: A proof of concept. International journal of

biological macromolecules. https://doi.org/10.1016/j.ijbiomac.2016.11.118

Noda, I., Green, P.R., Satkowski, M.M., Schechtman, L.A., 2005. Preparation and properties of a

novel class of polyhydroxyalkanoate copolymers. Biomacromolecules 6, 580–586.

https://doi.org/10.1021/bm049472m

Noda, I., Lindsey, S.B., Caraway, D., 2010. NodaxTM Class PHA Copolymers: Their Properties

and Applications, in: Chen, G.G.-Q. (Ed.), Plastics from Bacteria. Springer Berlin

Heidelberg, Berlin, Heidelberg, pp. 237–255.

59

Nuñez Oreza, L.A., Alvarez, A.F., Arias-Olguín, I.I., Torres Larios, A., Georgellis, D., 2012. The

ArcB Leucine Zipper Domain Is Required for Proper ArcB Signaling. PLoS One 7.

https://doi.org/10.1371/journal.pone.0038187

Nunn, W.D., 1986. A molecular view of fatty acid catabolism in Escherichia coli. Microbiological

reviews 50, 179.

O’Hagan, D., 2008. Understanding organofluorine chemistry. An introduction to the C–F bond.

Chem. Soc. Rev. 37, 308–319. https://doi.org/10.1039/B711844A

Okamura, E., Tomita, T., Sawa, R., Nishiyama, M., Kuzuyama, T., 2010. Unprecedented

acetoacetyl-coenzyme A synthesizing enzyme of the thiolase superfamily involved in the

mevalonate pathway. Proc. Natl. Acad. Sci. U.S.A. 107, 11265–11270.

https://doi.org/10.1073/pnas.1000532107

Oshima, T., Aiba, H., Masuda, Y., Kanaya, S., Sugiura, M., Wanner, B.L., Mori, H., Mizuno, T.,

2002. Transcriptome analysis of all two-component regulatory system mutants of

Escherichia coli K-12. Molecular microbiology 46, 281–291.

Overath, P., Pauli, G., Schairer, H.U., 1969. Fatty acid degradation in Escherichia coli. An

inducible acyl-CoA synthetase, the mapping of old-mutations, and the isolation of

regulatory mutants. Eur. J. Biochem. 7, 559–574.

Patten, C.L., Kirchhof, M.G., Schertzberg, M.R., Morton, R.A., Schellhorn, H.E., 2004.

Microarray analysis of RpoS-mediated gene expression in Escherichia coli K-12. Mol.

Genet. Genomics 272, 580–591. https://doi.org/10.1007/s00438-004-1089-2

Peña-Sandoval, G.R., Georgellis, D., 2010. The ArcB Sensor Kinase of Escherichia coli

Autophosphorylates by an Intramolecular Reaction. J Bacteriol 192, 1735–1739.

https://doi.org/10.1128/JB.01401-09

60

Peoples, O.P., Sinskey, A.J., 1989a. Poly-beta-hydroxybutyrate biosynthesis in Alcaligenes

eutrophus H16. Characterization of the genes encoding beta-ketothiolase and acetoacetyl-

CoA reductase. J. Biol. Chem. 264, 15293–15297.

Peoples, O.P., Sinskey, A.J., 1989b. Poly-beta-hydroxybutyrate (PHB) biosynthesis in Alcaligenes

eutrophus H16. Identification and characterization of the PHB polymerase gene (phbC). J.

Biol. Chem. 264, 15298–15303.

Perutz, R.N., Braun, T., 2007. 1.26 - Transition Metal-mediated C–F Bond Activation, in: Mingos,

D.M.P., Crabtree, R.H. (Eds.), Comprehensive Organometallic Chemistry III. Elsevier,

Oxford, pp. 725–758. https://doi.org/10.1016/B0-08-045047-4/00028-5

Pigeon, X., Bergeron, M., Barabé, F., Dubé, P., Frost, H.N., Paquin, J.-F., 2010. Activation of

allylic C-F bonds: palladium-catalyzed allylic amination of 3,3-difluoropropenes. Angew.

Chem. Int. Ed. Engl. 49, 1123–1127. https://doi.org/10.1002/anie.200904747

Pinto, A., Ciesla, J.H., Palucci, A., Sutliff, B.P., Nomura, C.T., 2016. Chemically Intractable No

More: In Vivo Incorporation of “Click”-Ready Fatty Acids into Poly-[(R)-3-

hydroxyalkanoates] in Escherichia coli. ACS Macro Letters 5, 215–219.

https://doi.org/10.1021/acsmacrolett.5b00823

Pipkorn, R., Waldeck, W., Didinger, B., Koch, M., Mueller, G., Wiessler, M., Braun, K., 2009.

Inverse-electron-demand Diels-Alder reaction as a highly efficient chemoselective ligation

procedure: synthesis and function of a BioShuttle for temozolomide transport into prostate

cancer cells. J. Pept. Sci. 15, 235–241. https://doi.org/10.1002/psc.1108

Plunkett, R.J., 1941. Tetrafluoroethylene polymers. US2230654A.

Poletto, F.S., Fiel, L.A., Donida, B., Ré, M.I., Guterres, S.S., Pohlmann, A.R., 2008. Controlling

the size of poly(hydroxybutyrate-co-hydroxyvalerate) nanoparticles prepared by

61

emulsification–diffusion technique using ethanol as surface agent. Colloids and Surfaces

A: Physicochemical and Engineering Aspects 324, 105–112.

https://doi.org/10.1016/j.colsurfa.2008.04.003

Preusting, H., Nijenhuis, A., Witholt, B., 1990. Physical characteristics of poly(3-

hydroxyalkanoates) and poly(3-hydroxyalkenoates) produced by Pseudomonas

oleovorans grown on aliphatic hydrocarbons. Macromolecules 23, 4220–4224.

https://doi.org/10.1021/ma00221a007

Purser, S., Moore, P.R., Swallow, S., Gouverneur, V., 2008. Fluorine in medicinal chemistry.

Chem. Soc. Rev. 37, 320–330. https://doi.org/10.1039/B610213C

Rahman, M., Hasan, M.R., Oba, T., Shimizu, K., 2006. Effect of rpoS gene knockout on the

metabolism of Escherichia coli during exponential growth phase and early stationary phase

based on gene expressions, enzyme activities and intracellular metabolite concentrations.

Biotechnol. Bioeng. 94, 585–595. https://doi.org/10.1002/bit.20858

Raman, N., Black, P.N., DiRusso, C.C., 1997. Characterization of the fatty acid-responsive

transcription factor FadR. Biochemical and genetic analyses of the native conformation

and functional domains. J. Biol. Chem. 272, 30645–30650.

Rand, J.M., Gordon, G.C., Mehrer, C.R., Pfleger, B.F., 2017. Genome sequence and analysis of

Escherichia coli production strain LS5218. Metab Eng Commun 5, 78–83.

https://doi.org/10.1016/j.meteno.2017.10.001

Riesenberg, D., Guthke, R., 1999. High-cell-density cultivation of microorganisms. Appl

Microbiol Biotechnol 51, 422–430. https://doi.org/10.1007/s002530051412

Robbins, J.E., 1988. A Proposed Pathway for Catabolism of Propionate in Methanogenic

Cocultures. Appl Environ Microbiol 54, 1300–1301.

62

Roeder, W., Somerville, R.L., 1979. Cloning the trpR gene. Mol. Gen. Genet. 176, 361–368.

Rostovtsev, V.V., Green, L.G., Fokin, V.V., Sharpless, K.B., 2002. A stepwise huisgen

cycloaddition process: copper(I)-catalyzed regioselective “ligation” of azides and terminal

alkynes. Angew. Chem. Int. Ed. Engl. 41, 2596–2599. https://doi.org/10.1002/1521-

3773(20020715)41:14<2596::AID-ANIE2596>3.0.CO;2-4

Saito, Y., Doi, Y., 1994. Microbial synthesis and properties of poly(3-hydroxybutyrate-co-4-

hydroxybutyrate) in Comamonas acidovorans. International Journal of Biological

Macromolecules 16, 99–104. https://doi.org/10.1016/0141-8130(94)90022-1

Salmon, K.A., Hung, S. -p., Steffen, N.R., Krupp, R., Baldi, P., Hatfield, G.W., Gunsalus, R.P.,

2005. Global Gene Expression Profiling in Escherichia coli K12: Effects of oxygen

availability and ArcA. Journal of Biological Chemistry 280, 15084–15096.

https://doi.org/10.1074/jbc.M414030200

Saxon, E., Bertozzi, C.R., 2000. Cell surface engineering by a modified Staudinger reaction.

Science 287, 2007–2010. https://doi.org/10.1126/science.287.5460.2007

Schaffitzel, C., Berg, M., Dimroth, P., Pos, K.M., 1998. Identification of an Na+-dependent

malonate transporter of Malonomonas rubra and its dependence on two separate genes. J.

Bacteriol. 180, 2689–2693.

Schilling, C.I., Jung, N., Biskup, M., Schepers, U., Bräse, S., 2011. Bioconjugation via azide-

Staudinger ligation: an overview. Chem Soc Rev 40, 4840–4871.

https://doi.org/10.1039/c0cs00123f

Scholz, C., Fuller, R.C., Lenz, R.W., 1994. Production of poly(β-hydroxyalkanoates) with β-

substituents containing terminal ester groups by Pseudomonas oleovorans. Macromol.

Chem. Phys. 195, 1405–1421. https://doi.org/10.1002/macp.1994.021950424

63

Seacat, A.M., Thomford, P.J., Hansen, K.J., Olsen, G.W., Case, M.T., Butenhoff, J.L., 2002.

Subchronic Toxicity Studies on Perfluorooctanesulfonate Potassium Salt in Cynomolgus

Monkeys. Toxicol Sci 68, 249–264. https://doi.org/10.1093/toxsci/68.1.249

Shi, M., Shoichet, M.S., 2008. Furan-functionalized co-polymers for targeted drug delivery:

characterization, self-assembly and drug encapsulation. J Biomater Sci Polym Ed 19,

1143–1157. https://doi.org/10.1163/156856208785540127

Silhavy, T.J., Pratt, L.A., 1995. Porin Regulon of Escherichia coli, in: Silhavy, T.J., Hoch, J.A.

(Eds.), Two-Component Signal Transduction. American Society of Microbiology, pp.

105–127.

Slater, S., Houmiel, K.L., Tran, M., Mitsky, T.A., Taylor, N.B., Padgette, S.R., Gruys, K.J., 1998.

Multiple beta-ketothiolases mediate poly(beta-hydroxyalkanoate) copolymer synthesis in

Ralstonia eutropha. J. Bacteriol. 180, 1979–1987.

Söhling, B., Gottschalk, G., 1996. Molecular analysis of the anaerobic succinate degradation

pathway in Clostridium kluyveri. J Bacteriol 178, 871–880.

Soper, T., Mandin, P., Majdalani, N., Gottesman, S., Woodson, S.A., 2010. Positive regulation by

small RNAs and the role of Hfq. Proceedings of the National Academy of Sciences 107,

9602–9607.

Sparks, J., Scholz, C., 2009. Evaluation of a cationic poly(β-hydroxyalkanoate) as a plasmid DNA

delivery system. Biomacromolecules 10, 1715–1719. https://doi.org/10.1021/bm900372x

Sparks, J., Scholz, C., 2008. Synthesis and Characterization of a Cationic Poly(β-

hydroxyalkanoate). Biomacromolecules 9, 2091–2096.

https://doi.org/10.1021/bm8005616

64

Spratt, S.K., Ginsburgh, C.L., Nunn, W.D., 1981. Isolation and genetic characterization of

Escherichia coli mutants defective in propionate metabolism. J Bacteriol 146, 1166–1169.

Steinbüchel, A., Valentin, H.E., 1995. Diversity of bacterial polyhydroxyalkanoic acids. FEMS

Microbiology Letters 128, 219–228. https://doi.org/10.1016/0378-1097(95)00125-O

Sudesh, K., Abe, H., Doi, Y., 2000. Synthesis, structure and properties of polyhydroxyalkanoates:

biological polyesters. Progress in Polymer Science 25, 1503–1555.

https://doi.org/10.1016/S0079-6700(00)00035-6

Takacs, M.L., Abbott, B.D., 2007. Activation of Mouse and Human Peroxisome Proliferator–

Activated Receptors (α, β/δ, γ) by Perfluorooctanoic Acid and Perfluorooctane Sulfonate.

Toxicol Sci 95, 108–117. https://doi.org/10.1093/toxsci/kfl135

Takagi, Y., Hashii, M., Maehara, A., Yamane, T., 1999. Biosynthesis of Polyhydroxyalkanoate

with a Thiophenoxy Side Group Obtained from Pseudomonas putida. Macromolecules 32,

8315–8318. https://doi.org/10.1021/ma981337c

Takase, K., Matsumoto, K., Taguchi, S., Doi, Y., 2004. Alteration of substrate chain-length

specificity of type II synthase for polyhydroxyalkanoate biosynthesis by in vitro evolution:

in vivo and in vitro enzyme assays. Biomacromolecules 5, 480–485.

Takase, K., Taguchi, S., Doi, Y., 2003. Enhanced synthesis of poly(3-hydroxybutyrate) in

recombinant Escherichia coli by means of error-prone PCR mutagenesis, saturation

mutagenesis, and in vitro recombination of the type II polyhydroxyalkanoate synthase

gene. J. Biochem. 133, 139–145.

Tanaka, J., Suzuki, S., Tokunaga, E., Haufe, G., Shibata, N., 2016. Asymmetric Desymmetrization

via Metal-Free C-F Bond Activation: Synthesis of 3,5-Diaryl-5-fluoromethyloxazolidin-2-

65

ones with Quaternary Carbon Centers. Angew. Chem. Int. Ed. Engl. 55, 9432–9436.

https://doi.org/10.1002/anie.201603210

Tappel, R.C., Kucharski, J.M., Mastroianni, J.M., Stipanovic, A.J., Nomura, C.T., 2012a.

Biosynthesis of Poly[(R)-3-hydroxyalkanoate] Copolymers with Controlled Repeating

Unit Compositions and Physical Properties. Biomacromolecules 13, 2964.

Tappel, R.C., Wang, Q., Nomura, C.T., 2012b. Precise control of repeating unit composition in

biodegradable poly(3-hydroxyalkanoate) polymers synthesized by Escherichia coli.

Journal of Bioscience and Bioengineering 113, 480–486.

https://doi.org/10.1016/j.jbiosc.2011.12.004

Theodorou, M.C., Panagiotidis, C.A., Panagiotidis, C.H., Pantazaki, A.A., Kyriakidis, D.A., 2006.

Involvement of the AtoS-AtoC signal transduction system in poly-(R)-3-hydroxybutyrate

biosynthesis in Escherichia coli. Biochim. Biophys. Acta 1760, 896–906.

https://doi.org/10.1016/j.bbagen.2006.01.020

Thompson, R.C., Olsen, Y., Mitchell, R.P., Davis, A., Rowland, S.J., John, A.W.G., McGonigle,

D., Russell, A.E., 2004. Lost at Sea: Where Is All the Plastic? Science 304, 838–838.

https://doi.org/10.1126/science.1094559

Thuronyi, B.W., Privalsky, T.M., Chang, M.C.Y., 2017. Engineered Fluorine Metabolism and

Fluoropolymer Production in Living Cells. Angewandte Chemie International Edition 56,

13637–13640. https://doi.org/10.1002/anie.201706696

Toro-Roman, A., Mack, T.R., Stock, A.M., 2005. Structural Analysis and Solution Studies of the

Activated Regulatory Domain of the Response Regulator ArcA: A Symmetric Dimer

Mediated by the α4-β5-α5 Face. J Mol Biol 349, 11–26.

https://doi.org/10.1016/j.jmb.2005.03.059

66

Tsuge, T., Taguchi, K., Doi, Y., others, 2003. Molecular characterization and properties of (R)-

specific enoyl-CoA hydratases from Pseudomonas aeruginosa: metabolic tools for

synthesis of polyhydroxyalkanoates via fatty acid β-oxidation. International journal of

biological macromolecules 31, 195–205.

Twigg, L.E., Socha, L.V., 2001. Defluorination of sodium monofluoroacetate by soil

microorganisms from central Australia. Soil Biology and Biochemistry 33, 227–234.

https://doi.org/10.1016/S0038-0717(00)00134-6

Valentin, H.E., Dennis, D., 1997. Production of poly(3-hydroxybutyrate-co-4-hydroxybutyrate) in

recombinant Escherichia coli grown on glucose. Journal of Biotechnology 58, 33–38.

https://doi.org/10.1016/S0168-1656(97)00127-2 van Aalten, D.M.F., DiRusso, C.C., Knudsen, J., Wierenga, R.K., 2000. Crystal structure of FadR,

a fatty acid-responsive transcription factor with a novel acyl coenzyme A-binding fold. The

EMBO Journal 19, 5167–5177. https://doi.org/10.1093/emboj/19.19.5167

Walker, J.R.L., Bong Chui Lien, 1981. Metabolism of fluoroacetate by a soil Pseudomonas sp.

and Fusarium solani. Soil Biology and Biochemistry 13, 231–235.

https://doi.org/10.1016/0038-0717(81)90026-2

Walker, M.C., Thuronyi, B.W., Charkoudian, L.K., Lowry, B., Khosla, C., Chang, M.C.Y., 2013.

Expanding the fluorine chemistry of living systems using engineered polyketide synthase

pathways. Science 341, 1089–1094. https://doi.org/10.1126/science.1242345

Wang, L.C., Morgan, L.K., Godakumbura, P., Kenney, L.J., Anand, G.S., 2012. The inner

membrane histidine kinase EnvZ senses osmolality via helix-coil transitions in the

cytoplasm. The EMBO Journal 31, 2648–2659. https://doi.org/10.1038/emboj.2012.99

67

Werber, F.X., Baptist, J.N., 1964. Poly-β-hydroxybutyric acid—a naturally occurring

thermoplastic material. Polymer Engineering & Science 4, 245–250.

https://doi.org/10.1002/pen.760040403

West, A.H., Stock, A.M., 2001. Histidine kinases and response regulator proteins in two-

component signaling systems. Trends in biochemical sciences 26, 369–376.

Winquist, A., Steenland, K., 2014. Perfluorooctanoic Acid Exposure and Thyroid Disease in

Community and Worker Cohorts. Epidemiology 25, 255–264.

https://doi.org/10.1097/EDE.0000000000000040

Wolf, C.J., Fenton, S.E., Schmid, J.E., Calafat, A.M., Kuklenyik, Z., Bryant, X.A., Thibodeaux,

J., Das, K.P., White, S.S., Lau, C.S., Abbott, B.D., 2007. Developmental Toxicity of

Perfluorooctanoic Acid in the CD-1 Mouse after Cross-Foster and Restricted Gestational

Exposures. Toxicol Sci 95, 462–473. https://doi.org/10.1093/toxsci/kfl159

Xu, Y., Heath, R.J., Li, Z., Rock, C.O., White, S.W., 2001. The FadR.DNA complex.

Transcriptional control of fatty acid metabolism in Escherichia coli. J. Biol. Chem. 276,

17373–17379. https://doi.org/10.1074/jbc.M100195200

Yang, S.Y., He, X.Y., Schulz, H., 1995. Glutamate 139 of the large alpha-subunit is the catalytic

base in the dehydration of both D- and L-3-hydroxyacyl-coenzyme A but not in the

isomerization of delta 3, delta 2-enoyl-coenzyme A catalyzed by the multienzyme complex

of fatty acid oxidation from Escherichia coli. Biochemistry 34, 6441–6447.

https://doi.org/10.1021/bi00019a025

Zhang, Y., Gao, R., Ye, H., Wang, Q., Feng, Y., 2014. A new glimpse of FadR-DNA crosstalk

revealed by deep dissection of the E. coli FadR regulatory protein. Protein & Cell 5, 928–

939. https://doi.org/10.1007/s13238-014-0107-3

68

69

Chapter 2: Enhancing poly(3-hydroxyalkanoate) production in Escherichia coli by the removal of the regulatory gene arcA

The work presented here was published in AMB Express (2016). Please see Appendix G for publication cover. My contributions include the gene deletion experiments, the growth and harvest of all bacterial cultures, the GC preparation and analysis, the polymer extraction and purification, the compositional verification using H-NMR, the molecular weight analysis using

GPC, the gene expression analysis, and the drafting of the manuscript in its entirety.

2.1 Abstract

Recombinant Escherichia coli is a desirable platform for the production of many biological compounds including poly(3-hydroxyalkanoates), a class of naturally occurring biodegradable polyesters with promising biomedical and material applications. Although the controlled production of desirable polymers is possible with the utilization of fatty acid feedstocks, a central challenge to this biosynthetic route is the improvement of the relatively low polymer yield, a necessary factor of decreasing the production costs. In this study we sought to address this challenge by deleting arcA and ompR, two global regulators with the capacity to inhibit the uptake and activation of exogenous fatty acids. We found that polymer yields in a ΔarcA mutant increased significantly with respect to the parental strain. In the parental strain, PHV yields were very low but improved 64-fold in the ΔarcA mutant (1.92 mg L-1 to 124 mg L-1) The a ΔarcA mutant also allowed for modest increases in some medium chain length polymer yields, while weight average molecular weights improved by approximately 1.5-fold to 12-fold depending on the fatty acid substrate utilized. These results were supported by an analysis of differential gene expression, which showed that the key genes (fadD, fadL, and fadE) encoding fatty acid degradation enzymes were all upregulated by 2-, 10-, and 31-fold in an ΔarcA mutant, respectively. Additionally, the

70 short chain length fatty acid uptake genes atoA, atoE and atoD were upregulated by 103-, 119-, and 303-fold respectively, though these values are somewhat inflated due to low expression in the parental strain. Overall, this study demonstrates that arcA is an important target to improve PHA production from fatty acids.

2.2 Introduction

Poly(3-hydroxyalkanoates), or PHAs, are a group of biodegradable polyesters produced by a variety of microorganisms as a form of carbon storage (Lu et al. 2009). These PHAs are typically classified as short chain-length (SCL) PHAs, which contain repeating units of 3 to 5 carbons, and medium chain-length (MCL) PHAs containing 6 to 14 carbons. The physical properties of PHAs are dependent on monomer composition; SCL PHAs are generally stiff and brittle while MCL

PHAs are more elastomeric, and co-polymerization of the two groups allows for great variability in material properties (Laycock et al. 2013). Previous studies have shown that MCL PHAs can be effectively produced in recombinant E. coli lacking the fatty acid degradation gene fadB when utilizing a related carbon source such as fatty acids, although the monomer composition of the resulting polymers was heterogenous and uncontrolled (Langenbach et al. 1997; Qi et al. 1997).

Recently, it was shown that the monomer identity can be precisely controlled in both PHA homo- and co-polymers synthesized by recombinant Escherichia coli strain LSBJ (Tappel et al.

2012a; Tappel et al. 2012b). This was accomplished by deleting both the fadB and fadJ genes in

E. coli LS5218, recombinantly co-expressing the phaJ4 gene from Pseudomonas putida KT2440 with the highly active and broad substrate utilizing phaC1(STQK) genes, and feeding in specific ratios of fatty acids for conversion to PHAs (Fig. 2.1). This system allowed for strict control of repeating unit composition which enables great control over the physical properties of PHA polymers produced using this system, unlocking the potential for tailoring PHA materials for click-

71 chemistry modifications (Levine et al. 2015; Levine et al. 2016; Pinto et al. 2016). Although these previous studies addressed control of the monomer composition and thus physical and chemical properties of PHA polymers, overall polymer yields were still relatively low and some fatty acid substrates had poor incorporation into either PHA homo- or copolymers. The previously defined system relied heavily on the deletion of the fadR gene in E. coli LS5218 for constitutive expression of the genes encoding enzymes from the β-oxidation pathway (Spratt et al. 1981). In a previous study, researchers demonstrated that the inhibition of β-oxidation intermediates using acrylic acid was an effective strategy for improving PHA biosynthesis, particularly in combination with a fadR deletion (Qi et al. 1998). In addition to FadR, there are three other transcriptional regulators, ArcA,

OmpR, and CRP-cAMP, that are known to inhibit the expression of genes involved in β-oxidation.

For this study, we focused on the regulators ArcA and OmpR because CRP-cAMP is known to act as a transcriptional activator of β-oxidation in the absence of glucose, and only exhibits repression when glucose is present (Fic et al. 2009). We hypothesized that removal of transcriptional regulators that inhibit expression of β-oxidation related genes would result in improved flux through fatty acid catabolic pathways to increase PHA polymer yields in our engineered system.

Therefore, in this study E. coli LSBJ was engineered to improve the biosynthesis of PHA from fatty acid substrates by removing the global regulatory genes arcA and ompR.

72

Figure 2.1: Biosynthesis of PHA in E. coli LSBJ utilizing short-chain-length (SCL) and medium- chain-length (MCL) fatty acids. The absence of fadB and fadJ in E. coli LSBJ in combination with the plasmid-borne recombinant enzymes PhaC1(STQK) and PhaJ4 establishes a linear pathway for the production of PHA polymers from free fatty acids. Extracellular fatty acids are transported across the outer membrane dependent on size; SCL and shorter MCL fatty acids can diffuse across the outer membrane, while longer MCL fatty acids can be transported by the long-chain fatty acid transporter FadL (Lepore et al. 2011). Inner membrane transport and activation is accomplished by the SCL-specific Ato system (AtoEAD) or the MCL-specific acyl-CoA synthetase FadD (Kameda and Nunn 1981; Theodorou et al. 2006). Acyl-CoA substrates are converted into enoyl- CoA by the acyl-CoA dehydrogenase enzyme FadE (Campbell and Cronan 2002), and are unable to proceed further through β-oxidation due to the absence of FadB and FadJ. The enoyl-CoA pool is then converted to (R)-3-hydroxyacyl-CoA by the R-specific enoyl-CoA hydratase PhaJ4 (Tsuge et al. 2003), and finally polymerized by the PHA synthase PhaC1(STQK) (Takase et al. 2003; Takase et al. 2004). This system allows for the biosynthesis of PHA polymers with tightly controlled repeating unit composition, as the number of carbons present in the fatty acid substrate is retained as the total number of carbons in each repeating unit.

73

The transcriptional regulator OmpR functions as a response regulator of the two component regulatory EnvZ/OmpR system, which exhibits control over the expression of outer membrane porins in response to osmolarity (Mizuno and Mizushima 1990). The sensor kinase EnvZ autophosphorylates in response to extracellular osmolarity and transphosphorylates OmpR

(OmpR-P), which binds to DNA and alters expression of genes within the OmpR regulon (Forst and Roberts 1994; Matsubara and Mizuno 1999). The most well-studied members of the OmpR regulon are OmpF and OmpC, outer membrane porins that control the diffusion of small hydrophilic molecules (Aiba and Mizuno 1990; Mizuno and Mizushima 1990; Silhavy and Pratt

1995). However, OmpR controls numerous other transporter genes including fadL (Table 2.1), as well as genes involved in amino acid metabolism and flagellar biosynthesis (Higashitani et al.

1993; Oshima et al. 2002).

The global transcriptional dual regulator ArcA functions as the response regulator of the two-component regulatory ArcAB system, which regulates the expression of genes involved in aerobic and anaerobic metabolism in response to oxygen availability (Iuchi and Lin 1988; Lynch and Lin 1996). During conditions of decreasing oxygen availability, ArcA is activated through a phosphorelay mechanism and binds to the consensus sequence 5’-wGTTAATTAw-3’ (w is A or

T) located in numerous genes, including several genes involved in β-oxidation shown in Table

2.1, and acts as either a repressor or activator (Iuchi and Lin 1992; Lynch and Lin 1996; Liu and

Wulf 2004; Cho et al. 2006; Peña-Sandoval and Georgellis 2010). It has previously been shown that the redox biochemistry and transcriptional regulation of an ΔarcA mutant strain is significantly altered during microaerobic growth conditions, and to a lesser degree during aerobic growth conditions (Oshima et al. 2002; Alexeeva et al. 2003; Shalel-Levanon et al. 2005). The work

74 presented here demonstrates for the first time an arcA deletion mutant combined with a fadR mutation to improve the biosynthesis of PHA polymers derived from fatty acid substrates.

Table 2.1: Regulation Targets of ArcA and OmpR

Target Gene Description Reference arcA fadL Long-chain fatty acid transporter, experimental evidence. Cho et al., 2006 fadD Acyl-CoA synthetase, experimental evidence. Cho et al., 2006 fadE Acyl-CoA dehydrogenase, experimental evidence. Cho et al., 2006 ompR fadL Long-chain fatty acid transporter, predicted. Higashitani et al., 1993

2.3 Methods

2.3.1 Materials

A complete list of strains, plasmids, and primers used for this study is shown in Table 2.2.

All strains were grown in Lennox Broth (LB; composition per liter: 10 g tryptone, 5 g yeast extract, and 5 g sodium chloride, pH 7.0) purchased from Difco, and the antibiotics kanamycin (50 mg L-

1) and ampicillin (100 mg L-1) were added to media throughout the experiment as appropriate. The fatty acids sodium butyrate (Alfa Aesar), pentanoic acid (Alfa Aesar), hexanoic acid (Alfa Aesar), heptanoic acid (Alfa Aesar), sodium octanoate (Sigma Aldrich), decanoic acid (Alfa Aesar), and dodecanoic acid (Acros Organics) were used as feed supplements for PHA production (12 mM), along with the surfactant Brij-35 (Fisher Scientific, 4.0 g L-1). Sodium phosphate dodecahydrate

(Acros Organics, 8 mM) was added when noted. Sodium hydroxide (5 M) was used to adjust the pH to 7.0 when necessary. Primers were ordered from Integrated DNA Technologies (IDT). ACS

HPLC-grade chloroform and methanol were used for gas chromatography (GC) sample preparation and polymer purification.

75

Table 2.2: Strains, plasmids, and primers.

Escherichia coli Relevant characteristics Source or reference LSBJ ΔfadB, ΔfadJ, atoC512 (Const), fadR601 Tappel et al., 2012 RSC02 ΔarcA LSBJ This study RSC04 ΔompR LSBJ This study RSC06 ΔarcA, ΔompR LSBJ This study

Plasmids pKD46 λ Red recombinase expression plasmid; expresses exo, β, and γ genes Datsenko and R from λ phage; ParaB promoter; araC; Amp ; temperature sensitive replicon Wanner, 2000 pKD13 Neomycin phosphotransferase flanked by FLP recombinase recognition Datsenko and targets, AmpR, KmR Wanner, 2000 pCP20 FLP recombinase expression plasmid, AmpR, temperature sensitive Datsenko and replicon Wanner, 2000 pBBR-C1J4SII pBBR1MCS-2 derivative ΔphaAB, phaJ4, phaC1 (STQK) Tappel et al. 2012

Primersa,b Sequence (5’ to 3’) pKD13.F.arcA ATGCAGACCCCGCACATTCTTATCGTTGAAGACGAGTTGGTAACACGCAAGTGTAGGCTGGAGCTGCTTC pKD13.R.arcA TTAATCTTCCAGATCACCGCAGAAGCGATAACCTTCACCGTGAATGGTGGATTCCGTGGATCCGTCGACC pKD13.F.ompR ATGCAAGAGAACTACAAGATTCTGGTGGTCGATGACGACATGCGCCTGCGGTGTAGGCTGGAGCTGCTTC pKD13.R.ompR TTAGAACATTACCTTATGACCGTACTGCTCAAGAATGCCTTTCACGCGTTATTCCGTGGATCCGTCGACC arcA.check.F/R GTTAATTTGCAGCATGCATCAGG / GACGATGAGTTACGTATCTGG ompR.check.F/R AAATTGTTGCGAACCTTTGG / GCAATAACGTACGGGCAAAT qAtoA.F/R GGTGCAGCCATGTTTGATAG / CGCGAGGTTTGCTTCTTC qAtoD.F/R ACTTGGCAACCTGACCTATC / GACCAGTTCATCTGGCTCTAC qAtoE.F/R ACTCGGTATCGCTTACCTTG / GCAGACCCGCAATCATAAAC qFadD.F/R TCTCCAGTCTGCATCTTTCC / CCATAGCCTTCCAGCAGATAC qFadE.F/R TTACCCGTCTGGATGAACTG / GACGGCTTTCTTCAGCTTTC qFadL.F/R GGGCGCTTCTATTACCTCTAA / TTTCAAGGTCGGTTGTACCC qRpoD.F/R GAGCAAGGCTATCTGACCTATG / GCCCATGTCGTTGATCATTTG a Underlined sequences are homologous to the gene to be deleted. b Forward and reverse primers are denoted with an F or R, respectively, and primers used for qPCR are denoted with a q.

2.3.2 Gene deletions

The deletion of the arcA and ompR genes was accomplished using the λ red recombinase

protocol, a commonly used method for nonpolar gene deletion, as previously described (Datsenko

and Wanner 2000; Tappel et al. 2012b). Briefly, knockout cassettes were generated using PCR

with gene-specific primers and the kanamycin resistance marker from pKD13 (Table 2.2). PCR

was performed using PrimeSTAR HS polymerase (Takara) following the manufacturers

76 recommended protocol. The λ red recombinase was expressed using plasmid pKD46, and knockout cassettes introduced by electroporation (1500V, 5ms; BTX ECM 399). Successful recombination was determined by antibiotic selection and loci screening using check primers

(Table 2.2). Antibiotic resistance was removed by the expression of FLP recombinase from the pCP20 plasmid, and successful deletions were confirmed by loss of antibiotic resistance and by

PCR using loci check primers (Table 2.2). Deletion mutants ΔarcA, ΔompR, and the double deletion ΔarcA ΔompR) were named RSC02, RSC04, and RSC06, respectively (Table 2.2).

2.3.3 PHA production

Protocols for PHA production and cell harvest were adapted from a previous study (Tappel et al. 2012b), with several modifications. For initial preliminary experiments, LSBJ, RSC02,

RSC04, and RSC06 were made chemically competent and transformed with pBBR-C1J4SII following standard procedures (Sambrook and Russell 2001), to express PHA synthase and enoyl-

CoA hydratase. Transformants were grown on LB-agar plates, and single colonies were used to inoculate separate 2 mL LB seed cultures, in triplicate for each strain. Seed cultures were grown for 16 h at 37°C and 200 rpm, and used to inoculate 100 mL of growth media in 500-mL baffled shake flasks (to final concentration of 0.5%). Growth media contained LB, Brij-35, decanoic acid, and kanamycin. Cultures were grown for 48 h at 30°C and 250 rpm on a rotary shaker, and were then harvested following previously published methods (Tappel et al. 2012b).

A more robust test of PHA homopolymer production was performed between LSBJ and

RSC02 using the methods described above, with two key differences. A variety of fatty acids were tested in the growth media: sodium butyrate, pentanoic, hexanoic, heptanoic, sodium octanoate, decanoic, and dodecanoic. In addition, sodium phosphate dodecahydrate was added to the growth media for these experiments, which acts as a buffer for the shorter chain fatty acids. This addition

77 also provides a significant source of phosphate to the media, so to keep growth conditions consistent, sodium phosphate dodecahydrate was added to every other experiment in this study.

2.3.4 GC analysis

The yields and repeating unit compositions of PHA polymers were determined using GC, as previously described (Braunegg et al. 1978; Tappel et al. 2012b). Briefly, dried cells (15-20 mg) were dissolved in 2 mL of sulfuric acid: methanol solution (15:85) and 2 mL of chloroform and heated at 100°C for 140 min in a 10 mL pressure vial (Kimax). The samples were cooled to room temperature followed by the addition of 1 mL of Nanopure filtered water, after which all samples were mixed by vortex. Aqueous and organic layers were allowed to separate for 20 min. The organic layer was passed through a 0.45 μm polytetrafluoroethylene (PTFE) syringe filter (Restek).

An aliquot of 500 μL of each filtered sample was mixed with 500 μL of methyl octanoate standard

(1 gL-1) in chloroform in a 2 mL GC vial. Samples were injected and separated using a GC 2010

Gas Chromatograph with an AOC-20i autoinjector with a flame ionization detector. Shimadzu’s

GCSolution software was used to analyze the data, and statistical significance of triplicate samples was determined using a two-tailed Student’s t-test with a 95% confidence interval (α = 0.05).

2.3.5 Polymer purification and molecular weight determination

PHA homopolymers were extracted from residual dried cell samples from the LSBJ and

RSC02 biosynthesis experiments by combining each set of triplicate samples into single 10 mL pressure vials (Kimax), adding 6 mL of chloroform, and incubating at 100 °C for 1 h. Each sample was filtered through a 0.45 μm PTFE syringe filter (Restek) into a 20-mL scintillation vial and rinsed twice with 2 mL aliquots of chloroform. Samples were concentrated to relative dryness using a rotary evaporator, and redissolved in 1 mL chloroform. Crude polymers were purified by non-solvent precipitation in cold methanol as described previously (Pinto et al. 2016), with several

78 modifications. Briefly, dissolved samples were added dropwise to 10 mL of ice-cold methanol

(4°C) with rapid stirring. The solution was centrifuged (3,452 x g, 30 min, 4°C) to pellet the PHA, decanted and washed with an additional 5 mL of methanol, and re-centrifuged. The supernatant was decanted, and the pellet dissolved in approximately 2 mL of chloroform to transfer to a scintillation vial. Samples were concentrated in a rotary evaporator and evaporated to dryness under high vacuum for 4 h.

The weight average (Mw) and number average (Mn) molecular weights for each sample were determined by gel permeation chromatography (GPC) as described previously (Pinto et al.

2016). Briefly, PHA solutions of approximately 1.0 g L-1 were prepared by dissolution in chloroform and passed through a syringe filter (0.45 μm PTFE). Samples were injected (50 μL) into a Shimadzu LC-20AD liquid chromatograph equipped with a Shimadzu SIL-20A autosampler, a Shimadzu CTO-20A column oven, and a Shimadzu RID-10A refractive index detector. Samples were passed through an 8 x 50 mm styrenedivinylbenzene (SDV) guard column

(5 μm particles; Polymer Standards Service) and an 8 x 300 mm SDV analytical column (5 μm particles; mixed bed porosity; max molecular weight 1E6 Da; Polymer Standards Service product sda0830051lim). The column oven was maintained at 40°C with a 1 mL min-1 mobile phase of chloroform. Molecular weight standards of polystyrene with a narrow polydispersity index were used for calibration. Shimadzu’s LCsolution software was used to analyze the data. GPC chromatograms are available as supplemental material (Appendix C).

2.3.6 Growth profiles of LSBJ and RSC02

An analysis of the growth profiles for LSBJ and RSC02 was performed under the PHA homopolymer biosynthesis conditions utilizing two separate substrates, sodium butyrate and decanoic acid. Media and growth conditions were identical to those in the PHA Production section.

79

A 1 mL aliquot was removed from each culture to measure the OD600 every hour using a spectrophotometer (Genesys 10S) until the stationary phase was observed. A 5 mL aliquot was removed from each culture at both 24 and 48 h to analyze PHA concentrations. The 5 mL samples were harvested and analyzed by GC as described above.

2.3.7 RNA extraction and qPCR analysis

RNA from E. coli LSBJ and RSC02 was isolated and purified as described previously

(Lundgren et al. 2013; Sarwar et al. 2016). To isolate RNA for real-time quantitative PCR (qPCR), each strain was grown in duplicate in 100 ml of growth media (as described above, with sodium phosphate dodecahydrate and decanoic acid) in 500-ml baffled shake flasks at 30°C and 250 rpm to an OD600 of ~0.6. Cultures were immediately stabilized by adding 1 ml of RNAprotect Bacteria reagent (Qiagen) to 0.5 ml of culture. Cells were then lysed with lysozyme and proteinase K as described in the manufacturer’s protocol. The total RNA was subsequently purified from the lysed cells with the RNeasy minikit (Qiagen) by using an on-column DNase digestion step. PCR and a

Bioanalyzer were used to check the RNA for DNA contamination, quality, and concentration.

The iScript cDNA synthesis kit (Bio-Rad) was used to generate cDNA from 1 μg of the purified RNA samples. A 10-fold dilution series of the pooled cDNA from the two duplicate RNA samples from LSBJ or RSC02 was used for the qPCR experiments. The qPCR experiments were performed in triplicate. The expression of several important genes for β-oxidation was normalized to rpoD, a housekeeping gene with stable expression during exponential growth (Jishage et al.

1996). Primers for qPCR were designed to produce ~100 bp amplicons of each of the following genes: atoA, atoD, atoE, fadD, fadE, fadL, and rpoD (Table 2.2). The qPCR mixtures contained

300 mM each primer, 10 μl of the 2× iQ SYBR green Supermix (Bio-Rad), 5 μl of diluted cDNA, and nuclease-free water to a total volume of 20 μl. qPCR was performed on the MiniOpticon

80 system (Bio-Rad) with the following conditions: 1 cycle of 95°C for 2 min and 40 cycles of 95°C for 15 s, 55°C for 30 s, and 72°C for 30 s. Window-of-linearity R2 values and amplification efficiency values ranged from 0.990 to 1.0 and 90.0% to >100%, respectively. The dilution series with the highest R2 values was used to calculate relative gene expression of RSC02 compared to

LSBJ using the Pfaffl method (Fleige et al. 2006). Amplification efficiencies and threshold cycle

(Cq) values were calculated using the program LinRegPCR (Ruijter et al. 2009).

2.4 Results

2.4.1 PHA homopolymer production

Our goal was to develop a strain capable of producing PHA polymers with controlled repeating unit compositions and increased yields. To achieve this, the mutant strains RSC02,

RSC04, and RSC06 were derived from E. coli LSBJ by the deletion of arcA, ompR, and a tandem arcA/ompR deletion, respectively. The amount of poly(3-hydroxydecanoate) (PHD) produced by these strains was then compared to E. coli LSBJ while expressing PhaJ4 and PhaC1(STQK) in a set of preliminary experiments. RSC02 produced significantly more PHD than other strains, with a yield of 0.353 g L-1 (Fig. 2.2). Strains RSC04 and RSC06 were not found to be significantly different from LSBJ (Fig. 2.2) (Table F1, Appendix F).

81

0.50 LSBJ 0.45 0.353* RSC02

0.40

) ) 1 - RSC04 0.35 RSC06 0.30 0.196 0.25 0.179 0.20 0.154 0.15

Polymer (gL Yield Polymer 0.10 0.05 0.00

Figure 2.2: Comparison of poly(3-hydroxydecanoate) (PHD) yield between LSBJ and mutant strains RSC02 (ΔarcA), RSC04 (ΔompR), and RSC06 (ΔarcA, ΔompR). The average yield achieved by RSC02 is significantly greater than that of LSBJ, while RSC04 and RSC06 are not significantly different from LSBJ. All values are averages of triplicate experiments plus or minus the standard deviation around those averages. An asterisk denotes a statistically significant difference compared to LSBJ (Student’s t-test, two-tailed, α = 0.05).

To further investigate the effect of the ΔarcA mutation on PHA polymer production, PHA homopolymer biosynthesis was characterized utilizing a variety of fatty acids. The fatty acids used for the production of PHA homopolymers were sodium butyrate, valeric acid, hexanoic acid, heptanoic acid, sodium octanoate, decanoic acid, and dodecanoic acid respectively. Analysis by

GC showed significant increases in the amount of polymer produced by E. coli RSC02 when

82 compared to E. coli LSBJ for nearly all fatty acid substrates, particularly for those with six or fewer carbons (Table 2.3). Overall the parental strain, LSBJ, produced very little short chain-length

PHAs, with yields of only 3.04 mg L-1 of poly(3-hydroxybutyrate) (PHB) and 1.92 mg L-1 of poly(3-hydroxyvalerate) (PHV), and only 44.8 mg L-1 of poly(3-hydroxyhexanoate) (PHHx), the shortest medium chain-length polymer (Table 2.3). We observed significant increases in the amount of PHA synthesized by RSC02 of approximately 3750%, 6360%, and 485% when cells utilized sodium butyrate, pentanoic acid, or hexanoic acid as substrates, respectively (Table 2.3).

For these shorter chain-length PHA polymers the identity was confirmed by 1H-NMR due to the large differences in production between strains (Fig. A1, Appendix A). Of those PHA polymers tested with greater than 6 carbons per repeating unit, only poly(3-hydroxyheptanoate) (PHHp) and

PHD yields were significantly different between strains, with an increase of approximately 61 and

115 percent observed for RSC02, respectively (Table 2.3).

Table 2.3: PHA yield comparison between LSBJ and RSC02.

PHA Concentrationa Percent Substrate Strain CDWa (g/L) PHAa (wt.%) (mg/L) Increaseb

Sodium butyrate LSBJ 0.80 ± 0.03 0.38 ± 0.05 3.04 ± 0.27 3750 RSC02 0.85 ± 0.03 13.7 ± 0.95 117 ± 11.1 * Pentanoic acid LSBJ 0.76 ± 0.05 0.23 ± 0.02 1.92 ± 0.42 6360 RSC02 1.14 ± 0.04* 11.1 ± 1.01 124 ± 7.25 * Hexanoic acid LSBJ 0.85 ± 0.01 5.47 ± 0.97 44.8 ± 6.13 485 RSC02 1.06 ± 0.01* 27.3 ± 4.28 262 ± 59.0 * Heptanoic acid LSBJ 0.93 ± 0.01 23.4 ± 0.48 198 ± 4.27 61 RSC02 1.02 ± 0.06 30.2 ± 2.28 319 ± 22.8 * Sodium octanoate LSBJ 1.22 ± 0.05 44.5 ± 8.68 543 ± 110 1.10 RSC02 1.03 ± 0.01* 54.1 ± 1.06 549 ± 12.5 Decanoic acid LSBJ 1.23 ± 0.05 29.3 ± 2.15 281 ± 152 115 RSC02 1.49 ± 0.02* 40.4 ± 1.30 603 ± 26.4 * Dodecanoic acid LSBJ 1.31 ± 0.17 23.5 ± 5.42 303 ± 38.4 7.26 RSC02 1.11 ± 0.03* 29.3 ± 5.21 325 ± 64.0 a All values are averages of biological triplicate experiments plus or minus the standard deviation about those averages. b Percent increase calculated as the increase in PHA concentration from RSC02 compared to LSBJ.

83

* Denotes statistically significant difference compared to LSBJ (Student’s t-test, two-tailed, α = 0.05).

2.4.2 Molecular weight comparison

To compare differences in physical properties of the polymers synthesized by E. coli LSBJ and RSC02, samples were analyzed by gel permeation chromatography (GPC) to determine the number average molecular weight (Mn), the weight average molecular weight (Mw), and the polydispersity (Mw/Mn) (Table 2.4). The polymers extracted from RSC02 had greater molecular weights than those from LSBJ for every polymer except poly(3-hydroxydodecanoate) (PHDD), which had a Mn of approximately 60 kDa and a Mw of approximately 172 kDa for both strains

(Table 2.4). In addition, the polymers extracted from RSC02 had a greater degree of polydispersity, again with the exception of PHDD (Table 2.4).

Table 2.4: PHA molecular weight data.

PHAa Strain Mw (kDa) Mn (kDa) Mw/Mn

PHB LSBJ ND ND ND RSC02 390 117 3.3 PHV LSBJ 18 16 1.2 RSC02 243 79 3.1 PHHx LSBJ 134 77 1.7 RSC02 408 171 2.4 PHHp LSBJ 219 106 2.1 RSC02 319 134 2.4 PHO LSBJ 157 73 2.1 RSC02 285 99 2.9 PHD LSBJ 145 50 2.9 RSC02 234 67 3.5 PHDD LSBJ 173 58 3.0 RSC02 172 60 2.9 a PHB, poly(3-hydroxybutyrate); PHV, poly(3-hydroxyvalerate); PHHx, poly(3- hydroxyhexanoate); PHHp, poly(3-hydroxyheptanoate); PHO, poly(3-hydroxyoctanoate); PHD, poly(3-hydroxydecanoate); PHDD, poly(3-hydroxydodecanoate); ND, not detected.

84

2.4.3 Growth profile of LSBJ and RSC02

We visually observed that RSC02 cultures appeared less optically dense than those of LSBJ during the first day of growth, which would be indicative of an increased lag time for this strain.

To quantify this observation, we investigated the growth profiles of LSBJ and RSC02 under PHA biosynthesis conditions supplemented with either sodium butyrate or decanoic acid. There was both an increase in lag time and a decrease in the growth rate of RSC02 relative to LSBJ for both substrates (Fig. 2.3). Regardless of strain, when the fatty acid substrate was sodium butyrate the lag phase was 1 – 2 h shorter and the stationary phase was reached more rapidly (Fig. 2.3). Despite the increased lag phase duration and slightly slower growth rate, RSC02 reached the same culture density as LSBJ when supplemented with sodium butyrate (OD600 ~3.5, 15 h), and reached a higher

OD600 when supplemented with decanoic acid of ~5.0 compared to ~3.6 in LSBJ by hour 17 (Fig.

2.3).

85

10

1

LSBJ-C4

600 RSC02-C4 OD 0.1 LSBJ-C10 RSC02-C10

0.01 0 4 8 12 16 Time (h) Figure 2.3: Growth profile of E. coli LSBJ and RSC02 during PHA-biosynthesis utilizing sodium butyrate (C4) or decanoic acid (C10). For both substrates, RSC02 displayed an extended lag time and slightly slower growth rate relative to LSBJ; however, RSC02 cultures reached approximately the same density as LSBJ upon reaching the stationary phase. For both strains, growth with sodium butyrate as the substrate caused a shorter lag phase than with decanoic acid substrate. All values are the averages of triplicate experiments plus or minus the standard deviation around those averages.

2.4.4 PHA yield at 24 vs 48 hours

To determine whether the increased lag phase in RSC02 had a negative effect on PHA yield earlier in the production cycle, small subsamples were removed from both LSBJ and RSC02 shake flasks after 24h and 48h during the growth profile experiments. Similar to the data shown in

Table 2.3, LSBJ produced very little PHB at either 24 or 48 hours, and there was very little change between the two time points (Fig. 2.4). However, LSBJ yields of PHD increased from 22.6% to

31.7% of cell dry weight, an increase of approximately 40% (Fig. 2.4). RSC02 produced significantly more polymer than LSBJ regardless of the time or fatty acid substrate; PHB yield

86 increased from 8.2% to 12.3% from 24 to 48 h, while PHD yield increased by ~20% between time points from 37.2% to 45.0% (Fig. 2.4).

60

45.0 50

37.2 40 31.7 LSBJ-C4 30 RSC02-C4 22.6 LSBJ-C10 20 RSC02-C10 12.3 Polymer (wt%) Yield Polymer 8.2 10 1.9 3.0 0 24 h 48 h

Figure 2.4: Comparison of PHA yield as a percentage of dry weight between LSBJ and RSC02 at 24 h and 48 h, utilizing either sodium butyrate (C4) or decanoic acid (C10). LSBJ saw insignificant changes in PHB production between 24 and 48 h, while RSC02 increased slightly from 8.2% to 12.3% of cell dry weight. PHD yield from LSBJ increased from 22.6% to 31.7% between 24 and 48 h, an increase of ~40%, while PHD yield from RSC02 increased from 37.2% to 45.0%, an increase of ~20%. Regardless of the time, RSC02 produced more polymer than LSBJ in all cases. All values are the averages of triplicate experiments plus or minus the standard deviation about those averages.

2.4.5 Relative gene expression of RSC02

To analyze the relative expression of fatty acid degradation genes, RNA was isolated from mid-exponential phase cultures of LSBJ and RSC02 and reverse transcribed into cDNA for qPCR.

An appropriate OD600 value (~ 0.6) for mid-exponential phase was determined using the previously defined growth profile (Fig. 2.3). All the genes analyzed were upregulated to some degree in

RSC02; SCL fatty acid uptake genes atoA, atoE and atoD were all upregulated to an extremely large degree (>100-fold), while fadD, fadE, and fadL were all upregulated to a lesser degree (2-,

87

10-, and 31-fold) (Fig. 2.5). It is important to note that the expression values reported for the ato system may be exaggerated due to the low number of transcripts observed for LSBJ (as observed by a late-cycle emergence of the fluorescence during qPCR analysis).

1000 303

119 103 atoD 100 31 atoE atoA 10 fadE 10

fadL Fold Change Fold 2 fadD

1 cDNA Transcript

Figure 2.5: Relative gene expression of RSC02 compared to LSBJ, normalized to rpoD and measured as fold changes. The expression of genes related to fatty acid degradation was significantly increased in the RSC02 strain during mid-exponential growth phase (OD600 of ~ 0.6). The SCL fatty acid degradation genes atoA, atoE, and atoD were massively upregulated by 103- fold, 119-fold, and 303-fold respectively. MCL fatty acid degradation genes fadD, fadL, and fadE were upregulated by 2-fold, 10-fold, and 31-fold respectively. Relative gene expression was calculated from qPCR fluorescence data using the LineRegPCR software (Ruijter et al. 2009) to calculate amplification efficiency and Cq values, and the Pfaffl method used to derive relative expression values with LSBJ as the calibrator (Fleige et al. 2006). All values are the averages of triplicate experiments plus or minus the standard deviation about those averages.

2.5 Discussion

The purpose of developing the mutant strains RSC02, RSC04, and RSC06 was to improve

PHA biosynthesis by removing regulatory genes known to interact with components of the PHA biosynthesis pathway. Based on the evidence seen in Fig. 2.2, the singular deletion of arcA conferred a significant increase in the production of PHD compared to the parental LSBJ. A

88 reasonable explanation for this observation is that ArcA becomes active in E. coli LSBJ during the growth conditions utilized for polymer production and inhibits the transcription of fadL, fadD, and fadE, which are known targets of ArcA regulation (Table 2.1). In an arcA deletion mutant this inhibition cannot occur, likely resulting in higher basal transcription of these fad genes. The deletion of ompR and the double deletion of ompR/arcA provided no benefit to PHD biosynthesis, and no statistically significant differences were observed between these mutants and LSBJ (Fig.

2.2). One possibility for this apparent lack of effect is that OmpR regulatory target FadL is not active towards decanoic acid, which has been demonstrated previously (Black 1990). However, the similarity between the RSC06 double deletion mutant and the RSC04 mutant suggests another possibility; the loss of regulation by OmpR may yield a mildly toxic phenotype that inhibits cell growth and metabolism. This possibility is supported by both the observation that both RSC04 and

RSC06 had significantly lower cell dry weights than either LSBJ and RSC02 (Table F1, Appendix

F), and by previous reporting from Oshima et al. that showed a marked growth deficiency in E. coli ΔompR mutants (2002). These findings provided motivation for further characterization of the

RSC02 mutant.

A variety of fatty acid substrates were used to characterize the differences in PHA homopolymer production between RSC02 and LSBJ. In nearly all cases, the deletion of arcA resulted in a significantly higher polymer yield (Table 2.3). Most surprisingly, the largest increases in polymer yield observed between RSC02 and LSBJ were for the two SCL polymers, PHB and

PHV, as well as the shortest MCL polymer, PHHx (Table 2.3). This result is interesting considering that there are fewer steps involved in SCL fatty acid metabolism that are regulated by

ArcA. Short chain fatty acids enter the β-oxidation cycle via enzymes derived from the ato operon, bypassing both FadL and FadD (Fig. 2.1). This operon has no known interaction with ArcA and

89 is instead regulated by the response regulator AtoC, which is present in both LSBJ and RSC02 as a constitutively expressed mutant that confers a high level of atoDAEB transcription (Spratt et al.

1981; Jenkins and Nunn 1987; Kyriakidis and Tiligada 2009). One possible explanation for this drastic difference in SCL PHA production is that secondary metabolite pools are different between the two strains; polyamines for example have been shown to have significant effects on the regulation of AtoC (Kyriakidis and Tiligada 2009), and previous studies have shown that aspects of polyamine metabolism are affected by the deletion of arcA (Partridge et al. 2006).

The differences in MCL PHA production between LSBJ and RSC02 are more varied than those observed for SCL polymers. No significant difference was found between strains when comparing the yields of poly(3-hydroxyoctanoate) (PHO) and PHDD; however, the yields of poly(3-hydroxyheptanoate) (PHHp) and PHD were significantly increased in RSC02 (Table 2.3).

The most reasonable explanation for this is that there are numerous differences in the substrate specificities of the enzymes involved in the PHA biosynthesis pathway. The long chain fatty acid transporter FadL is predominantly active on fatty acids containing 16 or more carbons, and no binding has been observed for decanoic acid, therefore it is unlikely that changes in fadL expression would account for the differences we observed (Black 1990). In contrast, the acyl-CoA synthetase FadD has high activity towards 12 and 10 carbon fatty acids and only low activity for

8 and 6 carbon fatty acids (Iram and Cronan 2006; Ford and Way 2015), while the recombinant enzymes PhaJ4 and PhaC1(STQK) each have their own well-documented substrate specificities that could contribute to the observed variation without any direct regulation by ArcA (Tsuge et al.

2003; Matsumoto et al. 2005; Sato et al. 2011). However, a complete understanding of fatty acid flux through this PHA biosynthesis pathway cannot be achieved with the current lack of information regarding the enzymatic activity of the acyl-CoA dehydrogenase FadE.

90

The molecular weight data showed a great degree of variability between PHAs, as well as between LSBJ and RSC02, with polydispersity values (Mw/Mn) from 1.2 to 3.3 (Table 2.4). The molecular weight of polymers produced by RSC02 were higher for all PHA polymers produced except for PHDD, and in general the polydispersity indices were higher for RSC02-derived polymers as well. Both the PHO and PHD molecular-number-average-molecular weights (Mn) were substantially lower for LSBJ than in previous studies, while the molecular weight and polydispersity for PHDD was found to be significantly higher than previously reported (Liu et al.

2011; Tappel et al. 2012b). A reliable comparison of the molecular weights observed in this study to other studies is difficult due to the variety of growth conditions and pathways employed.

However, a possible explanation for the differences observed between LSBJ and RSC02 is that increased basal expression of β-oxidation genes improves the supply of 3-hydroxy fatty acyl-CoA monomers to the PhaC1 (STQK) polymer synthase. The evidence summarized in Table 2.4 suggests that the molecular machinery in E. coli plays some role in determining the molecular weight of the polymers produced.

A comparison between the polymer yield data in Table 2.3 and the molecular weight data in Table 2.4 appears to support this explanation. The PHA homopolymers with the greatest increases in yield (PHB, PHV, and PHHx) similarly showed the greatest increases in molecular weight (with the partial exception of PHB, which was not extracted in a great enough quantity to detect by GPC). The PHHp and PHD obtained from RSC02, which also saw moderate improvement in terms of yield, was observed to have a higher molecular weight than that obtained from LSBJ, although to a lesser extent than the shorter chain length PHA homopolymers.

Interestingly, PHO produced by RSC02 had a higher molecular weight despite there being no significant difference in yield, while PHDD was not significantly different in either of those

91 measurements. While it appears that improving the utilization of fatty acids also improves PHA molecular weight using this biosynthetic platform, these minor discrepancies with regards to PHO reveal that this relationship is more complex than that.

Another possible explanation for this increase in molecular weights could be that less ethanol is produced by RSC02 than LSBJ. It is typical for bacterial cultures grown into stationary phase in shake flasks to reach some level of microaerobiosis, leading to the production of fermentative byproducts such as ethanol (Gupta and Rao 2003; Losen et al. 2004). It was previously reported that the supplementation of ethanol in cultures of recombinant E. coli led to a decrease in the molecular weight of PHB due to a chain transfer reaction from PhaC to ethanol

(Hiroe et al. 2013). In E. coli, the reversible enzyme AdhE is largely responsible for ethanol flux within the cell, allowing for both the biosynthesis and degradation of ethanol (Membrillo-

Hernandez et al. 2000). It has been previously shown that an ΔarcA mutant has an improved ethanol tolerance compared to wild type, and it has been hypothesized that this tolerance is derived from increased expression of adhE along with enzymes involved in the tricarboxylic acid cycle

(TCA), which could drive the breakdown of ethanol to acetyl-CoA (Goodarzi et al. 2010). There is also evidence in the literature showing that adhE expression is significantly higher in an E. coli

ΔarcA mutant than in wild type under a range of microaerobic conditions (Shalel-Levanon et al.

2005). It is therefore possible that endogenously produced ethanol is reducing PHA molecular weights via chain termination in LSBJ, and that this effect could be mitigated by an improved flux of ethanol back to acetyl-CoA. However, further investigation of this possibility is required as there is currently no direct evidence to support this explanation.

We investigated our observation that RSC02 had slower growth than LSBJ by recording hourly culture densities, and found that while RSC02 had a significantly longer lag phase, the

92 growth rate was nearly identical to LSBJ, with cultures reaching similar final cell densities regardless of which fatty acid substrate was provided (Fig. 2.3). These observations match the evidence found in the Keio collection of single-gene knockouts, which showed that arcA is nonessential, an ΔarcA mutant grows only slightly slower than wild type, and reaches only marginally lower culture densities (Baba et al. 2006). Despite the slightly slower growth observed with RSC02, the strain is capable of producing significantly more polymer than LSBJ even after only 24 hours of growth (Fig. 2.4). This offers an advantage when using this strain for a large- scale continuous fermentation, and shows that the lengthened lag phase does not significantly impede PHA biosynthesis.

Our analysis of the qPCR results shows a clear increase in the expression of each of the fatty acid degradation genes tested (Fig. 2.5). Although the improvements to fadD, fadL, and fadE were expected and support our hypothesis, the magnitude of the increase in the three ato genes came as a surprise. As we mentioned previously, ArcA is not known to directly regulate expression of the ato system, and both LSBJ and RSC02 harbor a constitutively expressed mutant atoC gene

(Spratt et al. 1981; Jenkins and Nunn 1987; Kyriakidis and Tiligada 2009). Comparing the drastic differences seen in SCL PHA yields between LSBJ and RSC02 provides strong evidence to suggest that the similarly drastic increases in gene expression are responsible for these effects

(Table 2.3). These results support the idea that the ΔarcA mutation indirectly effects the expression of the ato system, possibly mediated by an altered polyamine metabolism.

The differences between the relative expression of fadD, fadL and fadE also raise some interesting observations. For example, fadD was only modestly up-regulated (2-fold) and its protein product has a high activity towards both 10 and 12 carbon fatty acid substrates; however, while PHD biosynthesis was significantly improved PHDD biosynthesis was not (Table 2.3, Fig.

93

2.5). Similarly, fadL expression was significantly increased in RSC02 (10-fold) which does have limited binding affinity for 12 carbon fatty acids (Black 1990), and yet this does not appear to improve PHDD biosynthesis either. These results suggest that the binding affinity of FadE may be an important limiting factor in PHA biosynthesis using this system, considering that fadE was the most highly up-regulated of these three genes (31-fold) and yet only the yields of PHHx, PHHp, and PHD were significantly improved (Table 2.3, Fig. 2.5). While the binding affinity of FadE is not well-known, our results suggest that it may have low binding affinity towards octanoyl-CoA and dodecanoyl-CoA, although further investigation is needed to test that hypothesis.

One of the largest challenges still facing the PHA industry is the relatively low yield of polymer obtained, which contributes to the overall cost (Kaur 2015). The strain E. coli RSC02 that was developed in this study offers a significant improvement to the previously reported strain, E. coli LSBJ. The most significant improvements were seen in the biosynthesis of PHB, PHV, and

PHHx, with modest increases observed in PHHp and PHD. These results are supported by the increased expression of atoA, atoD, and atoE which correlates with improved PHB and PHV biosynthesis, and also the increased expression of fadD, fadL, and fadE which may contribute to the increased biosynthesis of PHHx, PHHp, and PHD.

2.6 References

Aiba, H., Mizuno, T., 1990. Phosphorylation of a bacterial activator protein, OmpR, by a protein

kinase, EnvZ, stimulates the transcription of the ompF and ompC genes in Escherichia coli.

FEBS Letters 261, 19–22. https://doi.org/10.1016/0014-5793(90)80626-T

Alexeeva, S., Hellingwerf, K.J., de Mattos, M.J.T., 2003. Requirement of ArcA for redox

regulation in Escherichia coli under microaerobic but not anaerobic or aerobic conditions.

Journal of Bacteriology 185, 204–209.

94

Baba, T., Ara, T., Hasegawa, M., Takai, Y., Okumura, Y., Baba, M., Datsenko, K.A., Tomita, M.,

Wanner, B.L., Mori, H., 2006. Construction of Escherichia coli K-12 in-frame, single-gene

knockout mutants: the Keio collection. Molecular Systems Biology 2.

https://doi.org/10.1038/msb4100050

Black, P.N., 1990. Characterization of FadL-specific fatty acid binding in Escherichia coli.

Biochimica et Biophysica Acta (BBA) - Lipids and Lipid Metabolism 1046, 97–105.

https://doi.org/10.1016/0005-2760(90)90099-J

Braunegg, G., Sonnleitner, B., Lafferty, R.M., 1978. A rapid gas chromatographic method for the

determination of poly-β-hydroxybutyric acid in microbial biomass. Eur. J. Appl. Microb.

Biotechnol. 6, 29 37. https://doi.org/10.1007/BF00500854

Campbell, J.W., Cronan, J.E., 2002. The enigmatic Escherichia coli fadE gene is yafH. J.

Bacteriol. 184, 3759–3764.

Cho, B.K., Knight, E.M., Palsson, B.O., 2006. Transcriptional regulation of the fad regulon genes

of Escherichia coli by ArcA. Microbiology 152, 2207–2219. https://doi.org/152/8/2207

[pii]

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

coli K-12 using PCR products. Proceedings of the National Academy of Sciences 97,

6640–6645.

Fic, E., Bonarek, P., Gorecki, A., Kedracka-Krok, S., Mikolajczak, J., Polit, A., Tworzydlo, M.,

Dziedzicka-Wasylewska, M., Wasylewski, Z., 2009. cAMP Receptor Protein from

Escherichia coli as a Model of Signal Transduction in Proteins - A Review. Journal of

Molecular Microbiology and Biotechnology 17, 1–11. https://doi.org/10.1159/000178014

95

Fleige, S., Walf, V., Huch, S., Prgomet, C., Sehm, J., Pfaffl, M.W., 2006. Comparison of relative

mRNA quantification models and the impact of RNA integrity in quantitative real-time

RT-PCR. Biotechnol Lett 28, 1601–1613. https://doi.org/10.1007/s10529-006-9127-2

Ford, T.J., Way, J.C., 2015. Enhancement of E. coli acyl-CoA synthetase FadD activity on medium

chain fatty acids. PeerJ 3, e1040. https://doi.org/10.7717/peerj.1040

Forst, S.A., Roberts, D.L., 1994. Signal transduction by the EnvZ-OmpR phosphotransfer system

in bacteria. Res. Microbiol. 145, 363–373.

Goodarzi, H., Bennett, B.D., Amini, S., Reaves, M.L., Hottes, A.K., Rabinowitz, J.D., Tavazoie,

S., 2010. Regulatory and metabolic rewiring during laboratory evolution of ethanol

tolerance in E. coli. Molecular Systems Biology 6. https://doi.org/10.1038/msb.2010.33

Gupta, A., Rao, G., 2003. A study of oxygen transfer in shake flasks using a non-invasive oxygen

sensor. Biotechnol. Bioeng. 84, 351–358. https://doi.org/10.1002/bit.10740

Higashitani, A., Nishimura, Y., Hara, H., Aiba, H., Mizuno, T., Horiuchi, K., 1993.

Osmoregulation of the fatty acid receptor gene fadL in Escherichia coli. Mol. Gen. Genet.

240, 339–347.

Hiroe, A., Hyakutake, M., Thomson, N.M., Sivaniah, E., Tsuge, T., 2013. Endogenous Ethanol

Affects Biopolyester Molecular Weight in Recombinant Escherichia coli. ACS Chemical

Biology 8, 2568–2576. https://doi.org/10.1021/cb400465p

Iram, S.H., Cronan, J.E., 2006. The β-Oxidation Systems of Escherichia coli and Salmonella

enterica Are Not Functionally Equivalent. Journal of Bacteriology 188, 599–608.

https://doi.org/10.1128/JB.188.2.599-608.2006

Iuchi, S., Lin, E.C., 1992. Purification and phosphorylation of the Arc regulatory components of

Escherichia coli. Journal of bacteriology 174, 5617–5623.

96

Iuchi, S., Lin, Ec., 1988. arcA (dye), a global regulatory gene in Escherichia coli mediating

repression of enzymes in aerobic pathways. Proceedings of the National Academy of

Sciences 85, 1888–1892.

Jenkins, L.S., Nunn, W.D., 1987. Genetic and molecular characterization of the genes involved in

short-chain fatty acid degradation in Escherichia coli: the ato system. Journal of

bacteriology 169, 42–52.

Jishage, M., Iwata, A., Ueda, S., Ishihama, A., 1996. Regulation of RNA polymerase sigma subunit

synthesis in Escherichia coli: intracellular levels of four species of sigma subunit under

various growth conditions. J. Bacteriol. 178, 5447–5451.

Kameda, K., Nunn, W.D., 1981. Purification and characterization of acyl coenzyme A synthetase

from Escherichia coli. J. Biol. Chem. 256, 5702–5707.

Kaur, G., 2015. Strategies for Large-scale Production of Polyhydroxyalkanoates. Chemical and

Biochemical Engineering Quarterly 29, 157–172.

https://doi.org/10.15255/CABEQ.2014.2255

Kyriakidis, D.A., Tiligada, E., 2009. Signal transduction and adaptive regulation through bacterial

two-component systems: the Escherichia coli AtoSC paradigm. Amino Acids 37, 443–458.

https://doi.org/10.1007/s00726-009-0241-z

Langenbach, S., Rehm, B.H.A., Steinbüchel, A., 1997. Functional expression of the PHA synthase

gene phaC1 from Pseudomonas aeruginosa in Escherichia coli results in poly(3-

hydroxyalkanoate) synthesis. FEMS Microbiology Letters 150, 303–309.

https://doi.org/10.1111/j.1574-6968.1997.tb10385.x

97

Laycock, B., Halley, P., Pratt, S., Werker, A., Lant, P., 2013. The chemomechanical properties of

microbial polyhydroxyalkanoates. Progress in Polymer Science 38, 536–583.

https://doi.org/10.1016/j.progpolymsci.2012.06.003

Lepore, B.W., Indic, M., Pham, H., Hearn, E.M., Patel, D.R., van den Berg, B., 2011. Ligand-

gated diffusion across the bacterial outer membrane. Proc Natl Acad Sci U S A 108, 10121–

10126. https://doi.org/10.1073/pnas.1018532108

Levine, A.C., Heberlig, G.W., Nomura, C.T., 2016. Use of thiol-ene click chemistry to modify

mechanical and thermal properties of polyhydroxyalkanoates (PHAs). International

Journal of Biological Macromolecules 83, 358–365.

https://doi.org/10.1016/j.ijbiomac.2015.11.048

Levine, A.C., Sparano, A., Twigg, F.F., Numata, K., Nomura, C.T., 2015. Influence of Cross-

Linking on the Physical Properties and Cytotoxicity of Polyhydroxyalkanoate (PHA)

Scaffolds for Tissue Engineering. ACS Biomaterials Science & Engineering 1, 567–576.

https://doi.org/10.1021/acsbiomaterials.5b00052

Liu, Q., Luo, G., Zhou, X.R., Chen, G.-Q., 2011. Biosynthesis of poly(3-hydroxydecanoate) and

3-hydroxydodecanoate dominating polyhydroxyalkanoates by β-oxidation pathway

inhibited Pseudomonas putida. Metabolic Engineering 13, 11–17.

https://doi.org/10.1016/j.ymben.2010.10.004

Liu, X., Wulf, P.D., 2004. Probing the ArcA-P Modulon of Escherichia coli by Whole Genome

Transcriptional Analysis and Sequence Recognition Profiling. J. Biol. Chem. 279, 12588–

12597. https://doi.org/10.1074/jbc.M313454200

98

Losen, M., Frölich, B., Pohl, M., Büchs, J., 2004. Effect of Oxygen Limitation and Medium

Composition on Escherichia coli Fermentation in Shake-Flask Cultures. Biotechnol

Progress 20, 1062–1068. https://doi.org/10.1021/bp034282t

Lu, J., Tappel, R.C., Nomura, C.T., 2009. Mini-Review: Biosynthesis of Poly(hydroxyalkanoates).

Polymer Reviews 49, 226 248. https://doi.org/10.1080/15583720903048243

Lundgren, B.R., Thornton, W., Dornan, M.H., Villegas-Peñaranda, L.R., Boddy, C.N., Nomura,

C.T., 2013. Gene PA2449 Is Essential for Glycine Metabolism and Pyocyanin Biosynthesis

in Pseudomonas aeruginosa PAO1. J. Bacteriol. 195, 2087–2100.

https://doi.org/10.1128/JB.02205-12

Lynch, A.S., Lin, E.C., 1996. Transcriptional control mediated by the ArcA two-component

response regulator protein of Escherichia coli: characterization of DNA binding at target

promoters. Journal of bacteriology 178, 6238–6249.

Matsubara, M., Mizuno, T., 1999. EnvZ-independent phosphotransfer signaling pathway of the

OmpR-mediated osmoregulatory expression of OmpC and OmpF in Escherichia coli.

Biosci. Biotechnol. Biochem. 63, 408–414. https://doi.org/10.1271/bbb.63.408

Matsumoto, K., Takase, K., Aoki, E., Doi, Y., Taguchi, S., 2005. Synergistic Effects of Glu130Asp

Substitution in the Type II Polyhydroxyalkanoate (PHA) Synthase: Enhancement of PHA

Production and Alteration of Polymer Molecular Weight. Biomacromolecules 6, 99–104.

https://doi.org/10.1021/bm049650b

Membrillo-Hernandez, J., Echave, P., Cabiscol, E., Tamarit, J., Ros, J., Lin, E.C., 2000. Evolution

of the adhE gene product of Escherichia coli from a functional reductase to a

dehydrogenase. Genetic and biochemical studies of the mutant proteins. J. Biol. Chem.

275, 33869–33875. https://doi.org/10.1074/jbc.M005464200

99

Mizuno, T., Mizushima, S., 1990. Signal transduction and gene regulation through the

phosphorylation of two regulatory components: the molecular basis for the osmotic

regulation of the porin genes. Molecular Microbiology 4, 1077–1082.

https://doi.org/10.1111/j.1365-2958.1990.tb00681.x

Oshima, T., Aiba, H., Masuda, Y., Kanaya, S., Sugiura, M., Wanner, B.L., Mori, H., Mizuno, T.,

2002. Transcriptome analysis of all two-component regulatory system mutants of

Escherichia coli K-12. Molecular microbiology 46, 281–291.

Partridge, J.D., Scott, C., Tang, Y., Poole, R.K., Green, J., 2006. Escherichia coli Transcriptome

Dynamics during the Transition from Anaerobic to Aerobic Conditions. Journal of

Biological Chemistry 281, 27806–27815. https://doi.org/10.1074/jbc.M603450200

Peña-Sandoval, G.R., Georgellis, D., 2010. The ArcB Sensor Kinase of Escherichia coli

Autophosphorylates by an Intramolecular Reaction. J Bacteriol 192, 1735–1739.

https://doi.org/10.1128/JB.01401-09

Pinto, A., Ciesla, J.H., Palucci, A., Sutliff, B.P., Nomura, C.T., 2016. Chemically Intractable No

More: In Vivo Incorporation of “Click”-Ready Fatty Acids into Poly-[(R)-3-

hydroxyalkanoates] in Escherichia coli. ACS Macro Letters 5, 215–219.

https://doi.org/10.1021/acsmacrolett.5b00823

Qi, Q., Rehm, B.H.A., Steinbüchel, A., 1997. Synthesis of poly(3-hydroxyalkanoates) in

Escherichia coli expressing the PHA synthase gene phaC2 from Pseudomonas aeruginosa:

comparison of PhaC1 and PhaC2. FEMS Microbiology Letters 157, 155–162.

https://doi.org/10.1111/j.1574-6968.1997.tb12767.x

Qi, Q., Steinbüchel, A., Rehm, B.H.A., 1998. Metabolic routing towards polyhydroxyalkanoic

acid synthesis in recombinant Escherichia coli (fadR): inhibition of fatty acid β-oxidation

100

by acrylic acid. FEMS Microbiology Letters 167, 89–94. https://doi.org/10.1111/j.1574-

6968.1998.tb13212.x

Ruijter, J.M., Ramakers, C., Hoogaars, W.M.H., Karlen, Y., Bakker, O., van den Hoff, M.J.B.,

Moorman, A.F.M., 2009. Amplification efficiency: linking baseline and bias in the analysis

of quantitative PCR data. Nucleic Acids Research 37, e45–e45.

https://doi.org/10.1093/nar/gkp045

Sambrook, J.J., Russell, D.W., 2001. Molecular cloning: A laboratory manual. Cold Spring Harbor

Laboratory Press, Cold Spring Harbor, NY.

Sarwar, Z., Lundgren, B.R., Grassa, M.T., Wang, M.X., Gribble, M., Moffat, J.F., Nomura, C.T.,

2016. GcsR, a TyrR-Like Enhancer-Binding Protein, Regulates Expression of the Glycine

Cleavage System in Pseudomonas aeruginosa PAO1. mSphere 1, e00020-16.

https://doi.org/10.1128/mSphere.00020-16

Sato, S., Kanazawa, H., Tsuge, T., 2011. Expression and characterization of (R)-specific enoyl

coenzyme A hydratases making a channeling route to polyhydroxyalkanoate biosynthesis

in Pseudomonas putida. Applied Microbiology and Biotechnology 90, 951–959.

https://doi.org/10.1007/s00253-011-3150-5

Shalel-Levanon, S., San, K.-Y., Bennett, G.N., 2005. Effect of ArcA and FNR on the expression

of genes related to the oxygen regulation and the glycolysis pathway in Escherichia coli

under microaerobic growth conditions. Biotechnology and Bioengineering 92, 147–159.

https://doi.org/10.1002/bit.20583

Silhavy, T.J., Pratt, L.A., 1995. Porin Regulon of Escherichia coli, in: Silhavy, T.J., Hoch, J.A.

(Eds.), Two-Component Signal Transduction. American Society of Microbiology, pp.

105–127.

101

Spratt, S.K., Ginsburgh, C.L., Nunn, W.D., 1981. Isolation and genetic characterization of

Escherichia coli mutants defective in propionate metabolism. J Bacteriol 146, 1166–1169.

Takase, K., Matsumoto, K., Taguchi, S., Doi, Y., 2004. Alteration of substrate chain-length

specificity of type II synthase for polyhydroxyalkanoate biosynthesis by in vitro evolution:

in vivo and in vitro enzyme assays. Biomacromolecules 5, 480–485.

Takase, K., Taguchi, S., Doi, Y., 2003. Enhanced synthesis of poly(3-hydroxybutyrate) in

recombinant Escherichia coli by means of error-prone PCR mutagenesis, saturation

mutagenesis, and in vitro recombination of the type II polyhydroxyalkanoate synthase

gene. J. Biochem. 133, 139–145.

Tappel, R.C., Kucharski, J.M., Mastroianni, J.M., Stipanovic, A.J., Nomura, C.T., 2012a.

Biosynthesis of Poly[(R)-3-hydroxyalkanoate] Copolymers with Controlled Repeating

Unit Compositions and Physical Properties. Biomacromolecules 13, 2964.

Tappel, R.C., Wang, Q., Nomura, C.T., 2012b. Precise control of repeating unit composition in

biodegradable poly(3-hydroxyalkanoate) polymers synthesized by Escherichia coli.

Journal of Bioscience and Bioengineering 113, 480–486.

https://doi.org/10.1016/j.jbiosc.2011.12.004

Theodorou, M.C., Panagiotidis, C.A., Panagiotidis, C.H., Pantazaki, A.A., Kyriakidis, D.A., 2006.

Involvement of the AtoS-AtoC signal transduction system in poly-(R)-3-hydroxybutyrate

biosynthesis in Escherichia coli. Biochim. Biophys. Acta 1760, 896–906.

https://doi.org/10.1016/j.bbagen.2006.01.020

Tsuge, T., Taguchi, K., Doi, Y., others, 2003. Molecular characterization and properties of (R)-

specific enoyl-CoA hydratases from Pseudomonas aeruginosa: metabolic tools for

102 synthesis of polyhydroxyalkanoates via fatty acid β-oxidation. International journal of biological macromolecules 31, 195–205.

103

Chapter 3: Increased production of the value-added biopolymers poly(R-3- hydroxyalkanoate) and poly(γ-glutamic acid) from hydrolyzed paper recycling waste fines

The work presented here was published in Frontiers in Bioengineering and Biotechnology (2019).

Please see Appendix G for publication cover.

3.1 Abstract

Reject fines, a waste stream of short lignocellulosic fibers produced from paper linerboard recycling, are a cellulose-rich paper mill byproduct that can be hydrolyzed enzymatically into fermentable sugars. In this study, the use of hydrolyzed reject fines as a carbon source for bacterial biosynthesis of poly(R-3-hydroxyalkanoate) (PHA) and poly(γ-glutamic acid) (PGA) was investigated. Recombinant Escherichia coli harboring PHA biosynthesis genes were cultivated with purified sugars or crude hydrolysate to produce both poly(R-3-hydroxybutyrate) (PHB) homopolymer and medium chain length-containing copolymer (PHB-co-MCL). Wild-type

Bacillus licheniformis WX-02 were cultivated with crude hydrolysate to produce PGA. Both PHB and short chain-length-co-medium chain-length (SCL-co-MCL) PHA yields from crude hydrolysate were a 2-fold improvement over purified sugars, and the MCL monomer fraction was decreased slightly in copolymers produced from crude hydrolysate. PGA yield from crude hydrolysate was similarly increased 2-fold. The results suggest that sugars from hydrolyzed reject fines are a viable carbon source for PHA and PGA biosynthesis. The use of crude hydrolysate is not only possible but beneficial for biopolymer production, eliminating the need for costly separation and purification techniques. This study demonstrates the potential to divert a lignocellulosic waste stream into valuable biomaterials, mitigating the environmental impacts of solid waste disposal.

104

3.2 Introduction

Paper waste fines are cellulose fibers that have become too short for incorporation into paper products due to repeated recycling and must therefore be rejected from this process. These rejected waste fines make up a significant proportion of the waste stream from paper mills, which is becoming increasingly difficult to landfill due to transportation costs and legislation (Laurijssen et al., 2010; Villanueva and Wenzel, 2007). Since reject fines are predominantly composed of cellulose they can be readily hydrolyzed into monomeric sugars making them an attractive waste stream for the production of value-added products, including biofuels, platform chemicals, and biopolymers such as polyhydroxyalkanoates (Bhuwal et al., 2014; Galbe and Zacchi, 2002; Min et al., 2015; Wang et al., 2013; Zhang, 2008). Waste fines from the recycling of old corrugated cartons (OCC) are particularly valuable as a source of fermentable sugars, as they are typically high in cellulose and low in lignin and other inhibitory chemicals and minerals compared with deinked paper pulp from other waste streams (Min et al., 2015; Saini et al., 2020).

Polyhydroxyalkanoates (PHAs) are a diverse class of bacterially produced polyesters known for their biodegradability and biocompatibility, which occur naturally as a form of carbon storage (Lee, 1996; Lu et al., 2009). The physical characteristics of PHAs are dependent on both monomeric composition and molecular weight, and they range from stiff and brittle crystalline materials to flexible and elastomeric amorphous polymers (Laycock et al., 2013). Poly(R-3- hydroxybutyrate) (PHB) is the most abundant PHA from both natural and anthropogenic sources; however, this material is of limited use due to its high brittleness. The copolymerization of 3HB with other monomers, particularly those of medium chain-length (6-14 carbons), can improve toughness and elasticity for a more versatile material (Noda et al., 2005). The large-scale production and utilization of PHAs is mainly limited by the production cost, a large portion of

105 which stems from the cost of the feedstock and which is high relative to the production costs for petroleum-based plastics with similar properties. One way to address these costs is to examine alternative inexpensive feedstocks, which has sparked interest in lignocellulosic waste streams as a cheap carbon source.

Poly(γ-glutamic acid) (PGA) is another biopolymer that has generated interest as a renewable material for a number of applications. PGA is biosynthesized naturally by a variety of

Bacillus species, and is an edible, water-soluble, biodegradable, and anionic biopolymer (Bajaj and Singhal, 2011; Ogunleye et al., 2015). These properties make PGA suited for a variety of applications, including metal-ion binding and flocculation for wastewater treatment, composite materials for tissue engineering and drug delivery, and as a medicinal metal chelator for heavy metal removal (Inbaraj and Chen, 2012; Shih et al., 2001; Siao et al., 2009; Ye et al., 2006; Yokoi et al., 1996). Current research into improving PGA production for human use is focused on the metabolic engineering of various Bacillus species; notably, wild-type B. licheniformis WX-02 is capable of producing large amounts of PGA from glucose and glutamate and has been successfully engineered for enhanced biosynthesis (Cai et al., 2018, 2017).

The work presented in this study demonstrates the successful biosynthesis of both PHA and PGA biopolymers from crude hydrolyzed paper waste fines.

3.3 Methods

3.3.1 Hydrolysate from waste fines

A recycled liner board mill provided waste fines from the screw-press sludge. The analysis of the waste fines is provided in Min et al., 2018. The enzymatic hydrolysis procedure was conducted as described by Min and Ramarao (2017). Enzymatic hydrolysis was conducted at 50°

106

C and with commercially available CTec2 enzymes (Novozymes USA) at a substrate consistency

of 5%. All other details are described by Min et al., (2018) and Min and Ramarao (2017).

3.3.2 Media and cultivation

A complete list of strains and plasmids is shown in Table 3.1. All E. coli strains were

grown on LB-Lennox (LB; composition per liter: 10 g tryptone, 5 g yeast extract, and 5 g sodium

chloride, pH 7.0) purchased from Difco, with 15 g L-1 agar when needed. Glucose (Acros

Organics) and xylose (Sigma Aldrich) were supplemented as carbon sources when noted, as well

as purified or crude linerboard waste fines hydrolysate (Min et al., 2015). Bacillus licheniformis

WX-02 were maintained using nutrient broth no. 2 (Oxoid) media. For PGA biosynthesis, B.

licheniformis WX-02 was cultivated in the following PGA biosynthesis media (composition per

liter): 90 g glucose (or 10% crude hydrolysate, v/v), 40 g sodium glutamate, 10 g sodium citrate

trihydrate, 10 g sodium nitrate, 8 g ammonium chloride, 1 g potassium phosphate trihydrate, 1 g

magnesium sulfate heptahydrate, 1 g zinc sulfate heptahydrate, 1 g calcium chloride, 0.15 g

manganese sulfate monohydrate, pH 7.3 ± 0.1. E. coli strains were made chemically competent

and transformed by heat shock following standard procedures (Sambrook and Russell, 2001), and

selection was performed on LB agar (15 g L-) plates. The antibiotics kanamycin (50 mg L-1) and

ampicillin (100 mg L-1) were added to media for selection and plasmid retention as appropriate.

All liquid media cultures were cultivated using a rotary shaking incubator (New Brunswick

Scientific).

Table 3.1: Bacterial strains and plasmids. Strains/Plasmids Relevant Characteristics Reference BW25113 E. coli, Δ(araBAD)567, ΔlacZ4787(::rrnB3), λ-, rph-1, Δ(rhaBAD)568, Lessard et al. 1998, hsdR514 Datsenko and Wanner 2000 LSBJ E. coli LS5218, ΔfadB, ΔfadJ, atoC512 (Const), fadR601 Tappel et al. 2012 B. licheniformis Bacillus licheniformis WX-02, saline soil isolate. Wei et al. 2010 pBBRSTQKAB pBBR1MCS-2 derivative (lac promoter); phaC1(STQK), phaA, phaB Nomura et al. 2004 pTrcGK pTrc99a derivative (trc promoter); phaG, alkK Wang et al. 2012

107

PHA biosynthesis methods were adapted from previous studies (Wang et al., 2012; Tappel et al., 2014). Individual colonies of transformed bacteria harboring PHA biosynthesis genes

(pBBRSTQKAB or pBBRSTQKAB / pTrcGK) were used to inoculate separate 2 mL LB seed cultures, in triplicate for each strain. Seed cultures were grown for 16 h at 37 °C and 200 rpm and used to inoculate 100 mL of LB media in 500-mL baffled shake flasks (final concentration of

0.5%). Shake flasks were cultivated at 30 °C and 250 rpm rotary shaking for a total of 48 hrs. After reaching an OD600 of 1.0, cultures were induced with isopropyl-β-D-thiogalactoside (IPTG) at a final concentration of 1 mM. Carbon supplements (glucose, xylose, or pure hydrolysate at 20 g L-

1; or crude hydrolysate at 8% v/v) were added 3 hours post-IPTG induction. Cells were collected by centrifugation at 3716 × g for 15 minutes, washed once with 45 mL of 35% ethanol and once with 45 mL of water, and dried via lyophilization.

PGA biosynthesis methods were adapted from a previous study (Cai et al., 2018).

Individual colonies were used to inoculate separate 2 mL LB seed cultures, in triplicate, and incubated for 12 hours at 37 °C and 200 rpm. Seed cultures were used to inoculate 100 mL of PGA biosynthesis media in 500-mL baffled shake flasks (final concentration of 0.5%). Shake flasks were cultivated at 37 °C and 200 rpm for a total of 36 hours. To collect PGA, the pH of 10 mL of culture was adjusted to 2.0 using HCl (conc.) and centrifuged at 8000 × g for 10 minutes. The supernatant was collected and neutralized with NaOH (10M), then precipitated with 30 mL of

100% ethanol and mixed by vortex. The precipitate was collected by centrifugation at 8000 × g for 10 minutes, then dried via lyophilization to obtain a dry weight.

108

3.3.3 Analytical procedures

The sugar content of the crude hydrolysate was determined using 1H-NMR (Kiemle et al.,

2003). A calibration curve was generated from the integration of α anomeric proton peaks of pure glucose and xylose at known concentrations relative to glucosamine added as an internal standard

(5 g L-1). Crude hydrolysate was diluted 10-fold, doped with the glucosamine standard, and the sugar concentrations calculated from the calibration curve. Inductively coupled plasma optical emission spectroscopy (ICP-OES) was used to analyze abundance of metal ions (Optima 5300

DV). Hydrolysate was diluted 10-fold and introduced to the ICP-OES instrument at a flow rate of

1.5 mL min-1. Al, As, Ba, Cd, Co, Cu, Cr, Fe, Mo, P, Pb, S, and Zn were analyzed with an axial plasma view, while Ca, K, Mn, Mg, and Na were analyzed radially. Abundance was analyzed by measuring peak areas for each element compared to 4-point calibration curves of known standards.

The yields and repeating unit compositions of PHA polymers were determined using gas chromatography (GC), as previously described with slight modification (Braunegg et al., 1978;

Scheel et al., 2016). ACS HPLC-grade chloroform and methanol were used for GC sample preparation. Lyophilized cells (15-20 mg) were suspended in 2 mL of a 15% (v/v) sulfuric acid solution in methanol and 2 mL of chloroform and heated at 100 °C for 140 min in a 10 mL pressure vial (Kimax). The samples were cooled to room temperature, and 1 mL of Nanopure filtered water and 500 μL of methyl octanoate standard (0.25% v/v) in chloroform were added and mixed by vortex. Aqueous and organic layers were separated by centrifugation for 5 minutes at 700 rpm

(Marathon 6K, Fisher Scientific). The organic layer was passed through a 0.2 μm polytetrafluoroethylene (PTFE) filter using a vacuum manifold (Millex Samplicity) into 2 mL GC vials. Samples were injected and separated using a GC 2010 Gas Chromatograph with an AOC-

20i autoinjector and a flame ionization detector. Shimadzu’s GCSolution software was used to

109 analyze the data. Statistical analyses were performed using the Data Analysis Toolpak for

Microsoft Excel.

PGA was verified by 1H-NMR spectroscopy using a Bruker AVANCE III 600 MHz instrument. Spectra were processed with Bruker TopSpin v3.5pI2.

3.4 Results

3.4.1 Hydrolysate characterization

The concentration of glucose and xylose in the crude linerboard waste fines hydrolysate were calculated to be 98.5 g L-1 and 28.8 g L-1, respectively (Fig. A4, Appendix A). The metal ion composition of the crude hydrolysate was determined to be: Ca (5486 ppm), Na (1853 ppm), S

(21.0 ppm), Mg (20.0 ppm), K (13.0 ppm), Zn (4.70 ppm), Mn (4.44 ppm), Al (2.24 ppm), Ba

(1.56 ppm), P (1.12 ppm), Fe (0.30 ppm), Pb (0.24 ppm), Cu (0.19 ppm), Co (40 ppb), As (32 ppb), Mo (17 ppb), Cd (10 ppb), and Cr (7 ppb).

3.4.2 PHA analysis

PHB and SCL-co-MCL PHA were synthesized in recombinant BW25113 and LSBJ using several different carbon sources. The yield of PHB biosynthesized by BW25113 and LSBJ when crude hydrolysate was the carbon source was 6.88 g L-1 and 7.65 g L-1, respectively, which was greater than a 2-fold increase compared with any other pure carbon source analysed (Fig. 3.1A).

These results were determined to be statistically significant using a Student’s T-test (two-tailed, α

= 0.05) comparing yields from each pure carbon source to that of the hydrolysate. PHB yields from pure carbon sources were not significantly different between BW25113 and LSBJ, except for the mixed glucose/xylose which led to yields of 1.18 g L-1 and 3.13 g L-1, respectively.

The yields of SCL-co-MCL biosynthesized by BW25113 and LSBJ displayed similar trends to PHB homopolymer production; with crude hydrolysate, PHA yields were 6.82 g L-1 and

110

6.26 g L-1, respectively, which was slightly less than a 2-fold increase compared to the other carbon sources (Fig. 3.1B). These results were determined to be statistically significant using a Student’s

T-test (two-tailed, α = 0.05) comparing yields from each pure carbon source to that of the hydrolysate. In LSBJ, none of the SCL-co-MCL yields from pure carbons were significantly different. However, in BW25113 the utilization of xylose led to significantly higher yields (3.21 g

L-1) compared to glucose and the mixed glucose/xylose (Fig. 3.1B).

Figure 3.1: PHA produced by E. coli strains BW25113 and LSBJ harboring (A) pBBRSTQKAB or (B) pBBRSTQKAB and pTrcGK. Yields of PHB in g L-1 for each carbon source are shown for both strains in panel (A), while yields of PHB-co-MCL are shown in panel (B). Data shown are averages and standard deviations of 3 biological replicates. Hydrolysate denotes the crude hydrolysate (white bar), Glu/Xyl denotes purified hydrolysate sugars (grey striped bar), and Glucose and Xylose are store-bought pure sugars (black and solid grey bars, respectively).

The 3HB monomer fractions were substantially different between BW25113 and LSBJ, with the former strain producing copolymers with greater than 99.5% 3HB regardless of carbon source (Fig. 3.2). LSBJ produced copolymers with greater variability in their monomer content, though still dominated by 3HB monomers which ranged from 95.1% to 98.4% (Fig. 3.2). Due to

111 the high variation observed in copolymer composition from LSBJ, only PHAs produced from glucose and the mixed glucose/xylose were found to be significantly different between LSBJ and

BW25113 using a Student’s T-test (two-tailed, α = 0.05). The MCL monomer composition followed similar trends between carbon sources, but not between strains; polymers produced by

BW25113 contained no observable 3-hydroxyoctanoate (3HO) monomers, whereas 3HO constituted a significant percentage of the MCL fraction of PHA produced by LSBJ (Fig. 3.3). 3- hydroxydecanoate (3HD) was the other dominant MCL monomer in PHA produced by LSBJ, with the combined 3HO and 3HD fractions making up greater than 87% of the total MCL fraction. The

MCL monomer compositions were compared between each pure carbon source and the hydrolysate separately for both LSBJ and BW25113 using a two-factor ANOVA with replication

(α = 0.05), and each comparison was found to be statistically significant with the exception of xylose from the LSBJ strain.

Figure 3.2: 3HB monomer content of PHB-co-MCL produced by E. coli strains BW25113 and LSBJ harboring pBBRSTQKAB (phaC, phaA, phaB) and pTrcGK (phaG, alkK). Data shown are averages and standard deviations of 3 biological replicates. Hydrolysate denotes the crude hydrolysate (white bar), Glu/Xyl denotes purified hydrolysate sugars (grey striped bar), and Glucose and Xylose are store-bought pure sugars (black and solid grey bars, respectively).

112

3.4.3 PGA analysis

PGA was synthesized by Bacillus licheniformis WX-02 from either pure glucose or the crude hydrolysate. PGA yields from these two carbon supplements were 3.25 g L-1 and 6.46 g L-1

(standard deviations of 0.32 and 0.90), respectively, and were determined to be significantly different using a Student’s T-test (two-tailed, α = 0.05). The identity of the PGA was confirmed by 1H-NMR (Fig. A5, Appendix A). A minor impurity was observed in the 1H-NMR spectrum

(3.08–2.88 ppm), which was confirmed to be unconnected to the polymer backbone by COSY-

NMR (Fig. A6, Appendix A).

Figure 3.3: MCL monomer content of PHB-co-MCL produced by E. coli strains BW25113 and LSBJ harboring pBBRSTQKAB and pTrcGK. No 3HO monomers were detected in polymer produced by BW25113. 3HDD, 3-hydroxydodecanoate (grey striped bar); 3HD, 3- hydroxydecanoate (solid grey bar); 3HO, 3-hydroxyoctanoate (dotted bar); 3HHx, 3- hydroxyhexanoate (white bar).

3.5 Discussion

In this study there was a marked increase in both PHA and PGA biopolymer yields when substituting the crude hydrolysate for the pure sugar carbon sources. In comparison with previous

113 studies, PHB and SCL-co-MCL PHA yields from crude hydrolysate were significantly improved.

Using similar methodologies and an unrelated E. coli strain (JM109), Nomura et al. achieved a yield of 2.31 g L-1 of PHB from pure glucose (Nomura et al., 2004). Similarly, SCL-co-MCL polymers were previously produced with a yield of 3.49 g L-1 using the same growth conditions and E. coli LS5218, the parental strain of LSBJ (Tappel et al., 2014). For the PGA production experiment, our results are an improvement over early studies that use similar media and growth methods, where researchers observed a yield of 2.16 g L-1 (Wei et al., 2010). However, there have been much greater improvements in yield by optimizing media formulations and genetically modifying WX-02 to improve ATP supply, with recent yields as high as 43.81 g L-1 (Cai et al.,

2018, 2017; Wei et al., 2010).

Due to the complex composition of the crude hydrolysate, it is difficult to pinpoint and investigate specific hypotheses for the observed increase in biopolymer yield. One possible explanation for these results is the abundance of several important metal ions in the hydrolysate.

The most abundant metal in the hydrolysate was calcium which was present at physiologically relevant concentrations (~10 mM in hydrolysate supplemented shake flasks) (Holland et al., 1999).

Although the role of calcium in prokaryotes is not completely understood, it has been implicated in processes such as cell division, chemotaxis, and regulation of mechanosensitive ion channels

(Booth, 2014; Domínguez et al., 2015; Kung et al., 2010; Martins et al., 2011; Norris et al., 1996).

Interestingly, there is also evidence of non-proteinaceous PHB and polyphosphate acting as voltage-gated calcium channels in E. coli as a strategy to maintain calcium homeostasis, and the increased PHB yield observed in this study could be partially attributed to that (Das et al., 1997;

Reusch et al., 1995). However, the extraction and analysis methods in this study cannot differentiate between transmembrane PHB and intracellular granules, so this is merely speculation.

114

Mg2+ and K+ are both vital to bacterial survival, and the presence of these two ions in the hydrolysate may have enhanced bacterial growth (Epstein, 2003; Romani and Scarpa, 2000). A recent study found evidence that peptide-based media may be Mg2+-limited, and that E. coli grown on tryptone-based media supplemented with glucose were unable to completely utilize that glucose unless supplemented with Mg2+(Christensen et al., 2017). However, this would not explain the enhancement of PGA biosynthesis which was carried out in a defined medium with abundant

Mg2+. Transition metal cofactors such as Fe, Zn, Mn, and Cu are also beneficial for bacterial growth in modest concentrations, and the trace amounts present in the hydrolysate could have also contributed to improved biomass and biopolymer yields (Hood and Skaar, 2012). Although heavy metals such as Pb, As, and Cd are toxic to bacteria, they were not present in high enough concentrations to inhibit bacterial growth (Mitra et al., 1975; Neumann and Leimkühler, 2008;

Peng et al., 2007).

The concentration of glucose and other sugars in bacterial media is an important parameter to consider, and under-feeding or over-feeding can significantly alter metabolism (Stephanopoulos et al., 1998). Although 20 g L-1 of glucose is often used in PHA experiments, there is evidence that this concentration is above the optimum for sustained bacterial growth (Christensen et al., 2017;

Shang et al., 2003). In fed-batch cultures of Ralstonia eutropha, PHB yields were highest when glucose was maintained at 9 g L-1 and a decrease in biomass and PHB yield was observed as glucose concentration increased (Shang et al., 2003). Other studies have found evidence that carbon or nitrogen limitation can increase PHB yields in recombinant E. coli (Wang et al., 2009).

This may have contributed to the increased biopolymer yields observed in this study, as the crude hydrolysate was supplemented at lower sugar concentrations than pure sugars (7.30 g L-1 glucose and 2.13 g L-1 xylose for PHA experiments, 8.95 g L-1 glucose and 2.62 g L-1 xylose for PGA

115 experiments). Additionally, acetate was present in the hydrolysate at 1 g L-1 and has been previously shown to be utilized by E. coli LS5218 derivatives and incorporated into PHB

(Salamanca-Cardona et al., 2016, 2014). However, this amount of acetate would only contribute a maximum theoretical yield of 0.865 g L-1 PHB, which does not account for the total increase of

PHB in cells grown on hydrolysate as opposed to pure sugars (Fig. 3.1).

The copolymerization of PHB with MCL 3HA monomers often gives desirable material properties by reducing the crystallinity and brittleness of the resulting polymers, even with MCL fractions as low as 6 mol% (Matsusaki et al., 2000; Sudesh et al., 2000). The SCL-co-MCL PHAs produced by LSBJ in this study only incorporated a small amount of MCL monomers, and almost no MCL monomers were incorporated by BW25113 regardless of the carbon source (Fig. 3.2 &

3.3). There are several potential reasons for the low MCL incorporation; these MCL monomers are scavenged from fatty acid biosynthesis by the transacylase activity of PhaG, a putative 3- hydoxyacyl-ACP-CoA acyltransferase (Fiedler et al., 2000; Rehm et al., 2001). Free 3-hydroxy fatty acids can also be incorporated through the 3-hydroxyacyl-CoA ligase activity of AlkK (Wang et al., 2012). The reduced concentration of sugars from crude hydrolysate is likely to slow the growth rate of the cultures and shift metabolism away from fatty acid biosynthesis, resulting in less availability of 3-hydroxyacyl compounds (Li and Cronan, 1993; Takamura and Nomura,

1988). In E. coli LSBJ, the transcriptional regulator FadR, which both activates fatty acid biosynthesis and represses β-oxidation, is non-functioning (DiRusso et al., 1992; Iram and Cronan,

2005). This may explain why LSBJ was able to incorporate a greater percent of MCL monomers than BW25113, which has a fully functioning fatty acid biosynthesis pathway and is less likely to have large pools of 3-hydroxyacyl-ACP available.

116

In this study we have successfully biosynthesized PHB, SCL-co-MCL PHA, and PGA using sugars derived from the hydrolysis of linerboard recycling waste fines. Although sugars obtained from this hydrolysis can be purified and do indeed result in PHA biosynthesis, this can be a costly and time-consuming process and generates a low-concentration sugar solution that must be concentrated for bacterial fermentation. The work here has shown that the crude waste fines hydrolysate is a good source of lignocellulosic sugars for value-added biopolymer production.

3.6 References

Bajaj, I., Singhal, R., 2011. Poly (glutamic acid) – An emerging biopolymer of commercial

interest. Bioresource Technology 102, 5551–5561.

https://doi.org/10.1016/j.biortech.2011.02.047

Bhuwal, A.K., Singh, G., Aggarwal, N.K., Goyal, V., Yadav, A., 2014. Poly-β-hydroxybutyrate

production and management of cardboard industry effluent by new Bacillus sp. NA10.

Bioresour. Bioprocess. 1, 9. https://doi.org/10.1186/s40643-014-0009-5

Booth, I.R., 2014. Bacterial mechanosensitive channels: progress towards an understanding of

their roles in cell physiology. Curr Opin Microbiol 18, 16–22.

https://doi.org/10.1016/j.mib.2014.01.005

Braunegg, G., Sonnleitner, B., Lafferty, R.M., 1978. A rapid gas chromatographic method for the

determination of poly-β-hydroxybutyric acid in microbial biomass.

Eur.J.Appl.Microb.Biotechnol. 6, 29-37. https://doi.org/10.1007/BF00500854

Cai, D., Chen, Y., He, P., Wang, S., Mo, F., Li, X., Wang, Q., Nomura, C.T., Wen, Z., Ma, X.,

Chen, S., 2018. Enhanced production of poly-γ-glutamic acid by improving ATP supply in

metabolically engineered Bacillus licheniformis. Biotechnology and Bioengineering 115,

2541–2553. https://doi.org/10.1002/bit.26774

117

Cai, D., He, P., Lu, X., Zhu, C., Zhu, J., Zhan, Y., Wang, Q., Wen, Z., Chen, S., 2017. A novel

approach to improve poly-γ-glutamic acid production by NADPH Regeneration in Bacillus

licheniformis WX-02. Scientific Reports 7, 43404. https://doi.org/10.1038/srep43404

Christensen, D.G., Orr, J.S., Rao, C.V., Wolfe, A.J., 2017. Increasing growth yield and decreasing

acetylation in Escherichia coli by optimizing the carbon-to-magnesium ratio in peptide-

based media. Appl. Environ. Microbiol. 83, e03034-16.

https://doi.org/10.1128/AEM.03034-16

Das, S., Lengweiler, U.D., Seebach, D., Reusch, R.N., 1997. Proof for a nonproteinaceous

calcium-selective channel in Escherichia coli by total synthesis from (R)-3-

hydroxybutanoic acid and inorganic polyphosphate. PNAS 94, 9075–9079.

https://doi.org/10.1073/pnas.94.17.9075

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

coli K-12 using PCR products. PNAS 97, 6640–6645.

https://doi.org/10.1073/pnas.120163297

DiRusso, C.C., Heimert, T.L., Metzger, A.K., 1992. Characterization of FadR, a global

transcriptional regulator of fatty acid metabolism in Escherichia coli. Interaction with the

fadB promoter is prevented by long chain fatty acyl coenzyme A. J. Biol. Chem. 267, 8685–

8691.

Domínguez, D.C., Guragain, M., Patrauchan, M., 2015. Calcium binding proteins and calcium

signaling in prokaryotes. Cell Calcium 57, 151–165.

https://doi.org/10.1016/j.ceca.2014.12.006

Epstein, W., 2003. The roles and regulation of potassium in bacteria. Prog. Nucleic Acid Res. Mol.

Biol. 75, 293–320. https://doi.org/10.1016/s0079-6603(03)75008-9

118

Fiedler, S., Steinbüchel, A., Rehm, B.H.A., 2000. PhaG-mediated synthesis of poly(3-

hydroxyalkanoates) consisting of medium-chain-length constituents from nonrelated

carbon sources in recombinant Pseudomonas fragi. Appl. Environ. Microbiol. 66, 2117–

2124. https://doi.org/10.1128/AEM.66.5.2117-2124.2000

Galbe, M., Zacchi, G., 2002. A review of the production of ethanol from softwood. Appl Microbiol

Biotechnol 59, 618–628. https://doi.org/10.1007/s00253-002-1058-9

Holland, I.B., Jones, H.E., Campbell, A.K., Jacq, A., 1999. An assessment of the role of

intracellular free Ca2+ in E. coli. Biochimie 81, 901–907. https://doi.org/10.1016/s0300-

9084(99)00205-9

Hood, M.I., Skaar, E.P., 2012. Nutritional immunity: transition metals at the pathogen-host

interface. Nat Rev Microbiol 10. https://doi.org/10.1038/nrmicro2836

Inbaraj, B.S., Chen, B.-H., 2012. In vitro removal of toxic heavy metals by poly(γ-glutamic acid)-

coated superparamagnetic nanoparticles. Int J Nanomedicine 7, 4419–4432.

https://doi.org/10.2147/IJN.S34396

Iram, S.H., Cronan, J.E., 2005. Unexpected functional diversity among FadR fatty acid

transcriptional regulatory proteins. J. Biol. Chem. 280, 32148–32156.

https://doi.org/10.1074/jbc.M504054200

Kiemle, D.J., Stipanovic, A.J., Mayo, K.E., 2003. Proton NMR methods in the compositional

characterization of polysaccharides, in: Hemicelluloses: Science and Technology, ACS

Symposium Series. American Chemical Society, pp. 122–139. https://doi.org/10.1021/bk-

2004-0864.ch009

Kung, C., Martinac, B., Sukharev, S., 2010. Mechanosensitive channels in microbes. Annu. Rev.

Microbiol. 64, 313–329. https://doi.org/10.1146/annurev.micro.112408.134106

119

Laurijssen, J., Marsidi, M., Westenbroek, A., Worrell, E., Faaij, A., 2010. Paper and biomass for

energy?: The impact of paper recycling on energy and CO2 emissions. Resources,

Conservation and Recycling 54, 1208–1218.

https://doi.org/10.1016/j.resconrec.2010.03.016

Laycock, B., Halley, P., Pratt, S., Werker, A., Lant, P., 2013. The chemomechanical properties of

microbial polyhydroxyalkanoates. Progress in Polymer Science 38, 536–583.

https://doi.org/10.1016/j.progpolymsci.2012.06.003

Lee, S.Y., 1996. Bacterial polyhydroxyalkanoates. Biotechnology and Bioengineering 49, 1–14.

https://doi.org/10.1002/(SICI)1097-0290(19960105)49:1<1::AID-BIT1>3.0.CO;2-P

Lessard, I.A., Pratt, S.D., McCafferty, D.G., Bussiere, D.E., Hutchins, C., Wanner, B.L., Katz, L.,

Walsh, C.T., 1998. Homologs of the vancomycin resistance D-Ala-D-Ala dipeptidase

VanX in Streptomyces toyocaensis, Escherichia coli and Synechocystis: attributes of

catalytic efficiency, stereoselectivity and regulation with implications for function. Chem.

Biol. 5, 489–504. https://doi.org/10.1016/s1074-5521(98)90005-9

Li, S.J., Cronan, J.E., 1993. Growth rate regulation of Escherichia coli acetyl coenzyme A

carboxylase, which catalyzes the first committed step of lipid biosynthesis. J. Bacteriol.

175, 332–340. https://doi.org/10.1128/jb.175.2.332-340.1993

Lu, J., Tappel, R.C., Nomura, C.T., 2009. Mini-Review: Biosynthesis of poly(hydroxyalkanoates).

Polymer Reviews 49, 226-248. https://doi.org/10.1080/15583720903048243

Martins, A., Machado, L., Costa, S., Cerca, P., Spengler, G., Viveiros, M., Amaral, L., 2011. Role

of calcium in the efflux system of Escherichia coli. International Journal of Antimicrobial

Agents 37, 410–414. https://doi.org/10.1016/j.ijantimicag.2011.01.010

120

Matsusaki, H., Abe, H., Doi, Y., 2000. Biosynthesis and properties of poly(3-hydroxybutyrate-co-

3-hydroxyalkanoates) by recombinant strains of Pseudomonas sp. 61-3.

Biomacromolecules 1, 17–22. https://doi.org/10.1021/bm9900040

Min, B.C., Bhayani, B.V., Jampana, V.S., Ramarao, B.V., 2015. Enhancement of the enzymatic

hydrolysis of fines from recycled paper mill waste rejects. Bioresources and Bioprocessing

2, 40. https://doi.org/10.1186/s40643-015-0068-2

Min, B.C., Jampana, S.N., Thomas, C.M., Ramarao, B.V. (2018). Study of buffer substitution

using inhibitory compound CaCO₃ in enzymatic hydrolysis of paper mill waste fines.

Journal of Korea TAPPI 50, 77–82. https://doi.org/10.7584/JKTAPPI.2018.04.50.2.77

Min, B.C., Ramarao, B.V. (2017). Mechanisms of the inhibition of enzymatic hydrolysis of waste

pulp fibers by calcium carbonate and the influence of nonionic surfactant for mitigation.

Bioprocess Biosyst. Eng. 40, 799–806. https://doi.org/10.1007/s00449-017-1745-7

Mitra, R.S., Gray, R.H., Chin, B., Bernstein, I.A., 1975. Molecular mechanisms of accommodation

in Escherichia coli to toxic levels of Cd2+. Journal of Bacteriology 121, 1180–1188.

Neumann, M., Leimkühler, S., 2008. Heavy metal ions inhibit molybdoenzyme activity by binding

to the dithiolene moiety of molybdopterin in Escherichia coli. FEBS J. 275, 5678–5689.

https://doi.org/10.1111/j.1742-4658.2008.06694.x

Noda, I., Bond, E.B., Green, P.R., Melik, D.H., Narasimhan, K., Schechtman, L.A., Satkowski,

M.M., 2005. Preparation, properties, and utilization of biobased biodegradable NodaxTM

copolymers, in: Polymer Biocatalysis and Biomaterials, ACS Symposium Series.

American Chemical Society, pp. 280–291. https://doi.org/10.1021/bk-2005-0900.ch019

Nomura, C.T., Tanaka, T., Gan, Z., Kuwabara, K., Abe, H., Takase, K., Taguchi, K., Doi, Y.,

2004. Effective enhancement of short-chain-length−medium-chain-length

121

polyhydroxyalkanoate copolymer production by coexpression of genetically engineered 3-

ketoacyl-acyl-carrier-protein synthase III (fabH) and polyhydroxyalkanoate synthesis

genes. Biomacromolecules 5, 1457–1464. https://doi.org/10.1021/bm049959v

Norris, V., Grant, S., Freestone, P., Canvin, J., Sheikh, F.N., Toth, I., Trinei, M., Modha, K.,

Norman, R.I., 1996. Calcium signalling in bacteria. Journal of Bacteriology 178, 3677–

3682. https://doi.org/10.1128/jb.178.13.3677-3682.1996

Ogunleye, A., Bhat, A., Irorere, V.U., Hill, D., Williams, C., Radecka, I., 2015. Poly-γ-glutamic

acid: production, properties and applications. Microbiology 161, 1–17.

https://doi.org/10.1099/mic.0.081448-0

Peng, L., Lifang, R., Hongyu, X., Xi, L., Chaocan, Z., 2007. Study on the toxic effect of lead(II)

ion on Escherichia coli. Biol Trace Elem Res 115, 195–202.

https://doi.org/10.1007/BF02686030

Rehm, B.H.A., Mitsky, T.A., Steinbüchel, A., 2001. Role of fatty acid de novo biosynthesis in

polyhydroxyalkanoic acid (PHA) and rhamnolipid synthesis by Pseudomonads:

establishment of the transacylase (PhaG)-mediated pathway for PHA biosynthesis in

Escherichia coli. Appl. Environ. Microbiol. 67, 3102–3109.

https://doi.org/10.1128/AEM.67.7.3102-3109.2001

Reusch, R.N., Huang, R., Bramble, L.L., 1995. Poly-3-hydroxybutyrate/polyphosphate complexes

form voltage-activated Ca2+ channels in the plasma membranes of Escherichia coli.

Biophys. J. 69, 754–766. https://doi.org/10.1016/S0006-3495(95)79958-1

Romani, A.M., Scarpa, A., 2000. Regulation of cellular magnesium. Front. Biosci. 5, D720-734.

https://doi.org/10.2741/romani

122

Saini, S., Chutani, P., Kumar, P., Sharma, K.K., 2020. Development of an eco-friendly deinking

process for the production of bioethanol using diverse hazardous paper wastes. Renewable

Energy 146, 2362–2373. https://doi.org/10.1016/j.renene.2019.08.087

Salamanca-Cardona, L., Scheel, R.A., Bergey, N.S., Stipanovic, A.J., Matsumoto, K., Taguchi, S.,

Nomura, C.T., 2016. Consolidated bioprocessing of poly(lactate-co-3-hydroxybutyrate)

from xylan as a sole feedstock by genetically-engineered Escherichia coli. Journal of

Bioscience and Bioengineering 122, 406–414.

https://doi.org/10.1016/j.jbiosc.2016.03.009

Salamanca-Cardona, L., Scheel, R.A., Lundgren, B.R., Stipanovic, A.J., Matsumoto, K., Taguchi,

S., Nomura, C.T., 2014. Deletion of the pflA gene in Escherichia coli LS5218 and its effects

on the production of polyhydroxyalkanoates using beechwood xylan as a feedstock.

Bioengineered 5, 284–287. https://doi.org/10.4161/bioe.29595

Sambrook, J.J., Russell, D.W., 2001. Molecular cloning: A laboratory manual. Cold Spring Harbor

Laboratory Press, Cold Spring Harbor, NY. https://doi.org/10.1086/394015

Scheel, R.A., Ji, L., Lundgren, B.R., Nomura, C.T., 2016. Enhancing poly(3-hydroxyalkanoate)

production in Escherichia coli by the removal of the regulatory gene arcA. AMB Express

6, 120. https://doi.org/10.1186/s13568-016-0291-z

Shang, L., Jiang, M., Chang, H.N., 2003. Poly(3-hydroxybutyrate) synthesis in fed-batch culture

of Ralstonia eutropha with phosphate limitation under different glucose concentrations.

Biotechnology Letters 25, 1415–1419. https://doi.org/10.1023/A:1025047410699

Shih, I.L., Van, Y.T., Yeh, L.C., Lin, H.G., Chang, Y.N., 2001. Production of a biopolymer

flocculant from Bacillus licheniformis and its flocculation properties. Bioresource

Technology 78, 267–272. https://doi.org/10.1016/S0960-8524(01)00027-X

123

Siao, F.Y., Lu, J.F., Wang, J.S., Inbaraj, B.S., Chen, B.H., 2009. In vitro binding of heavy metals

by an edible biopolymer poly(γ-glutamic acid). Journal of Agricultural and Food

Chemistry 57, 777–784. https://doi.org/10.1021/jf803006r

Stephanopoulos, G., Aristidou, A.A., Nielsen, J., 1998. Metabolic Engineering: Principles and

Methodologies, 1st edition. ed. Academic Press.

Sudesh, K., Abe, H., Doi, Y., 2000. Synthesis, structure and properties of polyhydroxyalkanoates:

biological polyesters. Progress in Polymer Science 25, 1503–1555.

https://doi.org/10.1016/S0079-6700(00)00035-6

Takamura, Y., Nomura, G., 1988. Changes in the intracellular concentration of acetyl-CoA and

malonyl-CoA in relation to the carbon and energy metabolism of Escherichia coli K12. J.

Gen. Microbiol. 134, 2249–2253. https://doi.org/10.1099/00221287-134-8-2249

Tappel, R.C., Kucharski, J.M., Mastroianni, J.M., Stipanovic, A.J., Nomura, C.T., 2012.

Biosynthesis of poly[(R)-3-hydroxyalkanoate] copolymers with controlled repeating unit

compositions and physical properties. Biomacromolecules 13, 2964-2972.

https://doi.org/10.1021/bm301043t

Tappel, R.C., Pan, W., Bergey, N.S., Wang, Q., Patterson, I.L., Ozumba, O.A., Matsumoto, K.,

Taguchi, S., Nomura, C.T., 2014. Engineering Escherichia coli for improved production

of short-chain-length-co-medium-chain-length poly[(R)-3-hydroxyalkanoate] (SCL-co-

MCL PHA) copolymers from renewable nonfatty acid feedstocks. ACS Sustainable

Chemistry & Engineering 2, 1879–1887. https://doi.org/10.1021/sc500217p

Villanueva, A., Wenzel, H., 2007. Paper waste – Recycling, incineration or landfilling? A review

of existing life cycle assessments. Waste Management, Life Cycle Assessment in Waste

Management 27, S29–S46. https://doi.org/10.1016/j.wasman.2007.02.019

124

Wang, L., Sharifzadeh, M., Templer, R., Murphy, R.J., 2013. Bioethanol production from various

waste papers: Economic feasibility and sensitivity analysis. Applied Energy 111, 1172–

1182. https://doi.org/10.1016/j.apenergy.2012.08.048

Wang, Q., Tappel, R.C., Zhu, C., Nomura, C.T., 2012. Development of a new strategy for

production of medium-chain-length polyhydroxyalkanoates by recombinant Escherichia

coli via inexpensive non-fatty acid feedstocks. Applied and Environmental Microbiology

78, 519–527. https://doi.org/10.1128/AEM.07020-11

Wang, Q., Yu, H., Xia, Y., Kang, Z., Qi, Q., 2009. Complete PHB mobilization in Escherichia

coli enhances the stress tolerance: a potential biotechnological application. Microbial Cell

Factories 8, 47. https://doi.org/10.1186/1475-2859-8-47

Wei, X., Ji, Z., Chen, S., 2010. Isolation of halotolerant Bacillus licheniformis WX-02 and

regulatory effects of sodium chloride on yield and molecular sizes of poly-γ-glutamic acid.

Applied Biochemistry and Biotechnology 160, 1332–1340.

https://doi.org/10.1007/s12010-009-8681-1

Ye, H., Jin, L., Hu, R., Yi, Z., Li, J., Wu, Y., Xi, X., Wu, Z., 2006. Poly(gamma,L-glutamic acid)-

cisplatin conjugate effectively inhibits human breast tumor xenografted in nude mice.

Biomaterials 27, 5958–5965. https://doi.org/10.1016/j.biomaterials.2006.08.016

Yokoi, H., Arima, T., Hirose, J., Hayashi, S., Takasaki, Y., 1996. Flocculation properties of poly(γ-

glutamic acid) produced by Bacillus subtilis. Journal of Fermentation and Bioengineering

82, 84–87. https://doi.org/10.1016/0922-338X(96)89461-X

Zhang, Y.-H.P., 2008. Reviving the carbohydrate economy via multi-product lignocellulose

biorefineries. J Ind Microbiol Biotechnol 35, 367–375. https://doi.org/10.1007/s10295-

007-0293-6

125

Chapter 4: Optimizing a fed-batch high-density fermentation process for medium chain- length poly(3-hydroxyalkanoates) in Escherichia coli

The work presented here is being prepared for submission to Frontiers in Bioengineering and

Biotechnology.

4.1 Abstract

Production of medium chain-length (MCL) poly(3-hydroxyalkanoates) (PHA) of homopolymeric composition via high-density fed-batch fermentation was investigated in β-oxidation deficient

Escherichia coli LSBJ. These homopolymeric PHAs exhibit attractive material properties such as increased flexibility and elasticity relative to PHB; however, these polymers are difficult to biosynthesize in native PHA-producing organisms, and there is a paucity of research towards developing high-density cultivation methods while retaining compositional control. In this study, we developed and optimized a fed-batch fermentation process in a stirred tank reactor, beginning with the biosynthesis of poly(3-hydroxydecanoate) (PHD) from decanoic acid by recombinant E. coli LSBJ with glucose as an energy source for growth. Bacteria were cultured in two phases, a biomass accumulation phase (37°C, pH 7.0) and a PHA biosynthesis phase (30°C, pH 8.0).

Through iterative optimizations of semi-defined media composition and glucose feed rate, 6.0 g of decanoic acid was converted to PHD with an 87.5% yield (4.54 g L-1). Stepwise increases in the amount of decanoic acid fed during the fermentation correlated with an increase in PHD, resulting in a final decanoic acid feed of 25 g converted to PHD at a yield of 89.4% (20.1 g L-1, 0.42 g L-1 h-1), at which point foaming became uncontrollable. Hexanoic acid, octanoic acid, 10-undecenoic acid, and 10-bromodecanoic acid were all individually supplemented at 20 g each and successfully polymerized with yields ranging from 66.8 to 99.0% (9.24 to 18.2 g L-1). Using this bioreactor strategy, co-fatty acid feeds of octanoic acid/decanoic acid and octanoic acid/10-azidodecanoic

126 acid (8:2 mol ratio each) resulted in the production of their respective copolymers at nearly the same ratio and at high yield, demonstrating that these methods can be used to control PHA copolymer composition.

4.2 Introduction

High-density fermentation of Escherichia coli has been studied for the last 50 years in an

-1 effort to achieve maximal cell densities (~200 g L dry cell weight), frequently to attain high volumetric productivity (g L-1 h-1) of a heterologously-expressed product (Shiloach and Fass,

2005). While this is often a product such as a protein or antibiotic, these techniques have also been employed to produce biopolymers such as poly(hydroxyalkanoates) (PHAs). PHAs are a broad class of bacterially-derived polyesters prized for their biodegradability and biocompatibility, which can be produced from renewable carbon sources such as sugars and lipids (Lu et al., 2009).

There are two dominant types of PHA characterized by the number of carbons present in each monomer, short chain-length (SCL) PHA containing 3-5 carbons, and medium chain-length

(MCL) PHA containing 6-14 carbons. The most commonly discussed PHAs in the literature are poly[(R)-3-hydroxybutyrate] (PHB) and SCL copolymers of PHB with 3-hydroxyvalerate, which exhibit stiff and brittle material properties with a high Young’s modulus (i.e. the measured stiffness of a material) (Lu et al., 2009), while MCL PHAs and SCL-co-MCL copolymers are the subject of more recent studies due to their more viscoelastic properties (Rai et al., 2011; Tappel et al.,

2012a).

Control over the monomeric composition of MCL PHAs can be achieved by culturing β- oxidation deficient (ΔfadBJ) E. coli LSBJ harboring the PHA biosynthesis genes phaJ4 and phaC1(STQK) in the presence of fatty acid substrates as previously described by our laboratory

(Tappel et al., 2012a, 2012b). Controlling the monomer composition confers control over the

127 material properties, which can be further customized by the incorporation of functionalized fatty acids that enable post-production chemical modifications (Levine et al., 2015; Pinto et al., 2016).

One of the main limitations of this MCL PHA biosynthesis platform is the low polymer yields obtained; the first reported MCL PHA biosynthesis in E. coli LSBJ reported yields from shake flask cultivations of approximately 0.26 – 0.4 g L-1 (Tappel et al., 2012b), which were later improved slightly by the deletion of the arcA transcriptional regulator to 0.26 – 0.6 g L-1 (Scheel et al., 2016), and most recently improved to 5.44 g L-1 by utilizing glucose as a co-substrate, doubling the culture duration, and heterologously expressing an acyl-CoA synthetase from

Pseudomonas putida (Mohd Fadzil et al., 2018).

Cultivating PHA-producing bacteria to a high-density in stirred tank reactors (STR, bioreactors) rather than shake flasks has been shown to greatly enhance volumetric productivity, and is a necessary step to scaling up production (Blunt et al., 2018; Koller, 2018; Lee et al., 1999).

However, the most common products created using these high-density fermentation techniques are

SCL PHAs and MCL ter- and quadripolymers with uncontrolled monomer composition, with only a few reports of successful MCL biosynthesis with controlled compositions. One notable exception is the biosynthesis of near-homopolymeric poly(3-hydroxydecanoate) (PHD) reported by Gao et al., who obtained a yield of 11.8 g L-1 (0.41 g L-1 h-1) from β-oxidation deficient Pseudomonas putida KT2440 (Gao et al., 2018).

The goal of this study was to optimize a fed-batch fermentation process for the high-density cultivation of recombinant E. coli LSBJ to produce MCL PHA polymers with controlled composition from fatty acid substrates. PHA biosynthesis from fatty acids is mostly decoupled from primary metabolism; exogenous fatty acids are imported and converted into 3-hydroxyacyl-

CoA via three steps of a modified β-oxidation pathway and are polymerized by the PHA synthase

128

PhaC1(STQK), while glucose catabolism drives cell growth and provides ATP and reducing equivalents (Fig. 5.1). Catabolite repression induced by the presence of glucose is one regulatory network expected to bridge the gap between primary metabolism and β-oxidation. Fatty acid metabolism is tightly regulated in E. coli and is subject to catabolite repression through the activity of the cyclic AMP receptor protein – cyclic AMP complex (CRP-cAMP), a global transcriptional regulator that activates transcription of fad genes in response to elevated cAMP levels in the absence of glucose (see Chapter 1 for more information on CRP); this activation is lost when glucose is present in sufficient concentrations (Deutscher, 2008; Feng and Cronan, 2012; Fic et al., 2009; Iram and Cronan, 2006).

The work presented in this study demonstrates the first successful biosynthesis of MCL

PHAs in E. coli LSBJ via high-density fermentation, and a subsequent process optimization to increase PHA yields and volumetric productivity.

129

Figure 4.1: Simultaneous glucose consumption and PHA biosynthesis in E. coli LSBJ. Glucose is metabolized solely for cell growth, generating ATP and reducing equivalents via glycolysis and the tricarboxylic acid (TCA) cycle. Extracellular fatty acids (R = 1-9 carbon) cross the outer membrane by diffusion and are simultaneously transported and activated by the acyl-CoA synthetase FadD (Kameda and Nunn, 1981; Lepore et al., 2011). Acyl-CoAs are converted to enoyl-CoAs by the acyl-CoA dehydrogenase FadE and are unable to proceed by native β-oxidation due to the absence of FadB and FadJ (Tappel et al., 2012b). Recombinant enzymes PhaJ4 and PhaC1(STQK), an R-specific enoyl-CoA hydratase and an engineered PHA synthase respectively, are conferred by the pBBR-C1J4SII plasmid and complete the PHA biosynthesis pathway (Table 4.1). MCL PHA biosynthesis from fatty acids is largely uncoupled from cell growth; however, both fadD and fadE expression are activated by the binding of the cyclic AMP receptor protein – cyclic AMP (CRP-cAMP) complex (Feng and Cronan, 2012). A high ATP concentration inversely lowers cAMP concentration, reducing the amount of complexed CRP and potentially decreasing the expression of fadD and fadE and preventing PHA biosynthesis. ATP shown in this figure is a 1 2 3 product of the intermediates NADH , GTP , and FADH2 (with the exception of some ATP generated by glycolytic enzymes). 130

4.3 Methods

4.3.1 Media and inoculum preparation

A complete list of strains, plasmids, and primers used for this study can be found in Table

4.1. Unless otherwise specified, all strains were maintained in LB-Lennox (LB; composition per liter: 10 g tryptone, 5 g yeast extract, and 5 g NaCl, pH 7.0) purchased from Difco, and the antibiotics kanamycin (50 mg L-1) and ampicillin (100 mg L-1) were added to media as necessary.

When necessary, Bacto agar (Difco) was added to media at a concentration of 15 g L-1 to make plates.

The fatty acids hexanoic acid (Sigma Aldrich), sodium octanoate (Sigma Aldrich), decanoic acid (Alfa Aesar), 10-undecenoic acid (Sigma Aldrich). 10-bromodecanoic acid (A.K.

Scientific), and 10-azidodecanoate (synthesized as described below) were supplemented as substrates for polymer production as noted. Fatty acid feed solutions for bioreactor fermentations were brought to pH 8.0 by the addition of NaOH (10 M) to form the respective conjugate base and either autoclaved (saturated fatty acids) or filter-sterilized (functionalized fatty acids) prior to addition to the bioreactor.

131

Table 4.1: Strains, plasmids, and primers.

Escherichia coli Relevant characteristics Source or reference

LSBJ ΔfadB, ΔfadJ, atoC512 (Const), fadR601 Tappel et al., 2012 RSC02 ΔarcA LSBJ Scheel et al., 2016 LSBJ Δcrp Δcrp LSBJ This study LSBJ CRP* crp*(I112L, T127I, A144T) LSBJ This study SM02 Δdld LSBJ CRP* This study SM04 Δdld, ΔarcA LSBJ CRP* This study SM06 Δdld, ΔarcA, Δaas LSBJ CRP* This study Plasmids pKD46 λ Red recombinase expression plasmid; expresses exo, β, and γ genes Datsenko and R from λ phage; ParaB promoter; araC; Ap ; temperature sensitive replicon Wanner, 2000 pKD13 Neomycin phosphotransferase flanked by FLP recombinase recognition Datsenko and targets, ApR, KmR Wanner, 2000 pCP20 FLP recombinase expression plasmid, ApR, temperature sensitive replicon Datsenko and Wanner, 2000 pJet1.2/blunt Linear blunt cloning vector with interruptible eco471R lethal gene, ApR Fisher Scientific pBBR-C1J4SII pBBR1MCS-2 derivative, phaJ4, phaC1 (STQK), KmR Tappel et al. 2012 pZS411 pTrc99a derivative, pct (Megasphaera elsdenii), ApR This study pBRL690 pBBR-C1J4SII derivative, tac promoter This study pRAS02 pJet1.2/blunt derivative, native crp (E. coli MG1655) This study pRAS05 pJet1.2/blunt derivative, engineered crp* This study pTrcDAEB pTrc99a derivative, atoDAEB operon (E. coli MG1655) This study Primersa,b,c Sequence (5’ to 3’) pKD13.F.arcA ATGCAGACCCCGCACATTCTTATCGTTGAAGACGAGTTGGTAACACGCAAGTGTAGGCTGGAGCTGCTTC pKD13.R.arcA TTAATCTTCCAGATCACCGCAGAAGCGATAACCTTCACCGTGAATGGTGGATTCCGTGGATCCGTCGACC pKD13.F.aas AATCTCCCTCCATTCGCTTTTACTGAATCAGAGCAAAGGGAGTTGGAATGTGTAGGCTGGAGCTGCTTC pKD13.R.aas ACCACAACGAAGTGTTAGTGTGCACTGACTCACTCATCGTGTTGTTCCGCATTCCGTGGATCCGTCGACC pKD13.F.crp ATGGTGCTTGGCAAACCGCAAACAGACCCGACTCTCGAATGGTTCTTGTCGTGTAGGCTGGAGCTGCTTC pKD13.R.crp TTAACGAGTGCCGTAAACGACGATGGTTTTACCGTGTGCGGAGATCAGGTATTCCGTGGATCCGTCGACC pKD13.F.dld ATGTCTTCCATGACAACAACTGATAATAAAGCCTTTTTGATGAACTTGCGTGTAGGCTGGAGCTGCTTC pKD13.R.dld TTACTCCACTTCCTGCCAGTTTTTCCGTTTACTGGTTTTACCGATCCCCGATTCCGGGGATCCGTCGACC crpM1.F/R AAACCCGGACcttCTGATGCGTT / ACCTGAATCAATTGGCGAAATTTTTTG crpM2.F/R TCTGCAAGTCattTCAGAGAAAGTGGG / CGACGCGCCATCTGTGCA crpM3.F/R GGGCCGCATTacaCAGACTCTGC / GTCACGTCGAGGAACGCC BL621.F/R TGATTAATTGTCAAGCCTGGGGTGCCTAATGA / TCGGCTCGTATAATGTGTGGAATTGTGAGCGG native.crp.F/R ATGGTGCTTGGCAAACCGCA / TAAACGAGTGCCGTAAACGAC atoDAEB.F GCAGAGCTCAAGGAGAAAATGAAAACAAAATTGATGACATTACAAGACGCCACC atoDAEB.R GCAGGTACCTTGATTAATTCAACCGTTCAATCACCATCGCAATTCC arcA.check.F/R GTTAATTTGCAGCATGCATCAGG / GACGATGAGTTACGTATCTGG aas.check.F/R AAATTGTTGCGAACCTTTGG / GCAATAACGTACGGGCAAAT crp.check.F/R ATCCAGAGACAGCGGCGTTA / AATGGCGCGCTACCAGGTAA dld.check.F/R CTCGCTGAATTGTTATACAAGGC / CACAGAACGATTAAGTGAATTCGG a Underlined sequences are homologous to the gene to be deleted. b Forward and reverse primers are denoted with an F or R, respectively. c Lowercase base-pairs indicate codons subjected to site-directed mutagenesis.

132

4.3.2 E. coli engineering and plasmid construction

The dld, arcA, and aas genes were deleted using homologous recombination via the λ Red recombinase protocol described by Datsenko and Wanner (2000). Knockout cassettes were generated by PCR (PrimeSTAR HS polymerase, Takara) using gene-specific primers to amplify the kanamycin resistance marker from pKD13, gel-purified, and concentrated by ethanol precipitation. Knockout cassettes were introduced to parental strains harboring pKD46 by electroporation (1500V, 5m; BTX ECM 399), recovered in 1 mL LB at 30 °C and 250 rpm for 1 h, and plated on LB agar supplemented with kanamycin at 37 °C for 24 h. For the dld knockout, colonies were patched onto minimal media agar plates (22 mM KH2PO4, 42 mM Na2HPO4,

-1 8.6 mM NaCl, 2.0 mM MgSO4, 18.9 mM NH4Cl, 15 g L agar, pH 7.0) that were supplemented with 25 mM D-lactate. Colonies that did not grow on the D-lactate minimal agar plates were considered putative dld::KmR mutants. Successful gene replacement for the arcA::KmR and aas::KmR mutants was confirmed by loci screening using PCR. The temperature-sensitive pKD46 plasmid was removed by growth for 12 hours at 42 °C. Antibiotic markers were removed by expression of FLP recombinase from the pCP20 plasmid, which was confirmed by the loss of kanamycin resistance and loci screening by PCR. The temperature-sensitive pCP20 plasmid was removed by growth for 12 hours at 42 °C. Successful mutants were prepared as 30% glycerol stocks and stored at -80 °C.

The crp* mutation, a group of mutations that exhibit reduced dependence on cAMP for

DNA binding and subsequent transcriptional regulation (Guidi-Rontani et al., 1981; Khankal et al., 2009), was introduced to E. coli LSBJ by first deleting the native crp gene and then replacing it with the mutated crp* gene. Native crp was removed following the same λ Red recombinase protocol described above. The crp* gene was generated from the native E. coli MG1655 crp gene

133 using the Q5® Site-Directed Mutagenesis Kit (New England Biolabs) following manufacturers recommendations. In brief, native crp was PCR amplified from MG1655 genomic DNA using the native.crp.F/native.crp.R primers and blunt ligated into pJet1.2/blunt to generate pRAS02. This plasmid was subjected to three successive rounds of site-directed mutagenesis using the crpM1, crpM2, and crpM3 primer sets to make the following substitutions, respectively: I112L, T127I, and A144T. The final crp* gene harboring all three mutations, designated pRAS05, was verified by Sanger sequencing (GENEWIZ, NJ).

The crp* gene was PCR amplified from pRAS05 using the native.crp.F/native.crp.R primers. The PCR mixture was treated with Dpn I and the desired 700 bp crp* product was gel purified. The purified crp* was concentrated via ethanol precipitation and subsequently electroporated into LSBJ ∆crp cells harboring the  Red plasmid pKD46. Cells were recovered in 1 mL of LB at 30 °C and 250 rpm for 1 h, and 0.5 mL of recovery culture was harvested by centrifugation and suspended in 0.2 mL of sterile water. Cell suspension (0.1 mL) was plated onto minimal medium agar consisting of 22 mM KH2PO4, 42 mM Na2HPO4, 8.6 mM NaCl, 2.0 mM

MgSO4, 20 mM pyruvate (carbon source), and 20 mM L-tryptophan (nitrogen source). This selection method was chosen because crp is required for growth on L-tryptophan as a nitrogen source in the presence of pyruvate as a carbon source (Isaacs et al., 1994). Plates were incubated at 37 °C for 48 h, and colonies visible after 48 h were patched onto LB agar plates and subsequently screened via Sanger sequencing to verify the presence of the crp* gene. The LSBJ CRP* mutant was prepared as a 30% glycerol stock and stored at -80 °C.

The pZS411 plasmid, harboring the pct gene necessary for lactyl-CoA formation from lactate, was constructed by excising the pct gene from the pTV118NpctC1AB(STQK) plasmid

(Taguchi et al., 2008) using 5` and 3` EcoRI sites and ligating into pTrc99a following standard

134 cloning procedures (Sambrook and Russell, 2001). The pTrcDAEB plasmid was constructed by amplifying the atoDAEB operon using the atoDAEB.F/R primer set and ligating this fragment into the blunt cloning vector pJet1.2/blunt following manufacturer’s specifications. The operon was then excised with 5` SacI and 3` KpnI, purified by gel electrophoresis, and ligated into pTrc99a.

The pBRL690 plasmid was constructed by mutating the lac promoter of pBBRC1J4SII into the tac promoter by site-directed mutagenesis of the -35 region using the BL621.F/R primer set and the Q5® site-directed mutagenesis kit.

4.3.3 Shake flask production

PHA production in shake flasks was performed as described in Chapter 2 (Scheel et al.,

2016). In brief, chemically competent E. coli LSBJ was transformed with MCL PHA biosynthesis plasmid pBBR-C1J4SII by heat-shock following standard procedures (Sambrook and Russell,

2001). Transformants were grown on LB-agar plates, and single colonies were used to inoculate three separate 2.5 mL LB seed cultures. Seed cultures were grown for 16 h at 37 °C and 200 rpm rotary shaking, and 500 μL were used to inoculate 100 mL of growth media (LB, 4.0 gL-1 Brij-35,

10 mM decanoic acid, 8 mM sodium phosphate dodecahydrate, kanamycin, pH 7.0) in 500-mL baffled shake flasks. Cultures were grown for 12, 24, 36, or 48 h at 30°C and 220 rpm, then harvested by centrifugation as previously described (Tappel et al., 2012b). Before harvesting, the pH of each culture was measured (Accumet® pH meter, Fisher Scientific) immediately after removal from the incubator.

4.3.4 Low-density fed-batch production

A summary of all bioreactor trials outlining the E. coli strains, plasmids, and conditions can be found in Table 4.2. All fed-batch fermentations were performed in a 2-L vessel with a

BioFlo 310 benchtop bioreactor system (New Brunswick Scientific) using conditions and media

135 adapted from Pfeifer et al. (2002). A schematic of this bioreactor assembly is shown in Fig. F5

(Appendix F). For initial experiments, the assembled bioreactor containing 750 mL of defined

-1 -1 -1 medium (F1 Salts; 0.4 g L (NH4)2SO4, 1.5 g L KH2PO4, and 4.35 g L K2HPO4, pH 7.0) was autoclaved at 121 °C for 35 min. Prior to inoculation, 7.5 g glucose, 1.5 g yeast extract, 0.12 g

-1 MgSO4, 1 mL of 0.05 g L kanamycin sulfate, and 0.75 mL of trace metal solution were added

-1 -1 -1 aseptically. Trace metal solution contained 5 g L NaCl, 1 g L ZnSO4 · 7H2O, 4 g L MnCl2 ·

-1 -1 -1 -1 4H2O, 4.75 g L FeCl3, 0.4 g L CuSO4 · 5H2O, 0.58 g L H3BO3, 0.5 g L NaMoO4 · 2H2O, and

-1 8 mL L concentrated H2SO4. The bioreactor was inoculated with 40 mL of bacterial seed culture, which was grown in LB for 16 h at 37 °C and 200 rpm rotary shaking, pelleted at 3,716 × g for 10 min, and resuspended in 2 mL PBS buffer before sterile transfer to the bioreactor vessel.

Fermentation was carried out in two phases, a biomass accumulation phase (Phase 1) and a polymer biosynthesis phase (Phase 2). The first phase of fermentation was performed at 37 °C with an initial agitation rate of 200 rpm and an aeration rate of 5.0 L min-1, and pH was maintained at 7.0 by automated addition of concentrated NH4OH (17M). Dissolved oxygen was maintained at

50% air saturation controlled by agitation and air (compressed atmospheric) sparging cascades

(200 to 950 rpm, then 5.0 to 20.0 L min-1 air flow). Foaming was controlled by automated addition of Antifoam B (J.T. Baker, 10% v/v unless otherwise noted). Once the initial glucose was consumed (as evidenced by a sudden increase in dissolved oxygen), a glucose feed solution (500

-1 -1 -1 mL, filter-sterilized) consisting of 430 g L glucose, 3.9 g L MgSO4, 110 g L (NH4)2SO4, and

50 mg L-1 kanamycin was supplied at a rate of 14.5 mL min-1 with a separate peristaltic pump.

OD600 measurements were made on 1 mL culture aliquots each hour using a Genesys 10S UV-Vis spectrophotometer (Thermo Scientific).

136

After 12 h of growth in Phase 1, Phase 2 was initiated by decreasing the fermentation temperature to 30 °C, increasing the pH setpoint to 8.0 (with the exception of preliminary experiments, Table 4.2), decreasing the dissolved oxygen setpoint to 30% air saturation, and beginning the addition of a fatty acid feed solution supplied at a rate of 5.5 mL min-1 (except where otherwise noted for preliminary experiments) with another peristaltic pump. For Trial 1, D-lactate

(11 g) was added alongside octanoic acid as a bulk supplement. Glucose feed addition during the second phase was either maintained at 14.5 mL min-1 or reduced to 9.7 mL min-1 where noted.

Culture samples (10 mL) were removed at 24 and 48 h and stored at -80 °C for later quantitative

PHA analysis. The remaining culture at 48 h was harvested as described below.

4.3.5 High-density fed-batch production

High-density fermentations were carried out in the same 2-L vessel, using media adapted from Lau et al. (2004). The F1 Salts medium was replaced with 750 mL of defined medium

-1 -1 -1 consisting of 0.4 g L (NH4)2SO4, 4.16 g L KH2PO4, and 11.95 g L K2HPO4. Prior to

-1 inoculation, 7.5 g glucose, 1.5 g yeast extract, 0.12 g MgSO4, 1 mL of 0.05 g L kanamycin sulfate, and 2.25 mL of trace metal solution were added aseptically. The amount of yeast extract was increased to 3.0 g for trial 15 and for each subsequent trial (Table 4.2). The amount of (NH4)2SO4 in the glucose feed was reduced to 6 g L-1, and 15 mL of trace metal solution was added to the feed as well. All other bioreactor conditions were the same as for the low-density production process.

-1 Fatty acid feed solutions initially contained 40 g L decanoic acid (6 g in 150 mL H2O, pH

8.0), and were raised in subsequent trials as shown in Table 4.2. To maintain a constant volume of the combined glucose and fatty acid feeds, as fatty acid feed volumes increased (due to a maximum decanoic acid solubility of 100 g L-1 at pH 8.0) the glucose feed volume was decreased by the same magnitude while holding the mass of feed components constant. To account for this

137 change in glucose feed concentration, the addition rate was reduced proportionally so that the g glucose h-1 was kept constant. Antifoam concentration was increased to 20% for Trials 18 and 24 to combat excessive foaming.

Table 4.2: Summary of Bioreactor Conditions

Trial1 Strain Plasmid Glucose Feed Fatty Acid Feed3; Fatty Acid Feed pH Rate (mL h-1) Total Mass (g) Rate (mL h-1) 1 SM02 pBBRC1J4SII/pZS411 14.5 C8; 0.8 Bulk 7 2 SM04 pBBRC1J4SII 14.5 C8; 6 10 7 3 SM06 pBBRC1J4SII 14.5 / 9.7 C8; 1.0 Bulk 7 4 SM06 pBBRC1J4SII 14.5 / 9.7 C8; 1.5 Bulk 7 5 SM06 pBBRC1J4SII 14.5 / 9.7 C8; 6 8.3 7 6 SM06 pBBRC1J4SII 14.5 / 9.7 C8; 6 5.5 7 74 SM06 pBRL690 14.5 / 9.7 C8; 6 5.5 7 84 LSBJ pBBRC1J4SII / 14.5 C10; 6 5.5 7 pTrcDAEB 9 LSBJ pBBRC1J4SII 14.5 C10; 6 5.5 7 / 8 10 RSC02 pBBRC1J4SII 14.5 C10; 6 5.5 7 / 8 11 LSBJ CRP* pBBRC1J4SII 14.5 C10; 6 5.5 7 / 8 12 LSBJ pBBRC1J4SII 14.5 / 9.7 C10; 6 5.5 7 / 8 13 LSBJ pBBRC1J4SII 14.5 / 9.7 C10; 6 5.5 7 / 8 14 LSBJ pBBRC1J4SII 14.5 / 9.7 C10; 12 5.5 7 / 8 152 LSBJ pBBRC1J4SII 14.5 / 9.7 C10; 12 5.5 7 / 8 16 LSBJ pBBRC1J4SII 14.5 / 9.7 C10; 15 5.5 7 / 8 17 LSBJ pBBRC1J4SII 12.9 / 8.7 C10; 20 6.5 7 / 8 18 LSBJ pBBRC1J4SII 11.4 / 7.7 C10; 25 7.5 7 / 8 19 LSBJ pBBRC1J4SII 12.9 / 8.7 C6; 20 6.5 7 / 8 20 LSBJ pBBRC1J4SII 12.9 / 8.7 C8; 20 6.5 7 / 8 21 LSBJ pBBRC1J4SII 12.9 / 8.7 C11Δ10; 20 6.5 7 / 8 22 LSBJ pBBRC1J4SII 12.9 / 8.7 C10Br; 20 6.5 7 / 8 23 LSBJ pBBRC1J4SII 12.9 / 8.7 C8/C10; 15.4/4 6.5 7 / 8 24 LSBJ pBBRC1J4SII 12.9 / 8.7 C8/C10N3; 15.4/4.95 6.5 7 / 8 1 Trials 1 – 8 are preliminary experiments which did not result in polymer production. Trials 1 – 12 were performed following low-density fermentation methods, while trials 13 – 24 were performed following high-density fermentation methods. Trials 9 – 17 were performed in duplicate as two consecutive bioreactor fermentations, with the exception of trial 14. 2 Yeast extract in culture medium was doubled to 3.0 g total for trial 15, and this amount was maintained for each subsequent trial. 3 Fatty acids are abbreviated as follows: octanoic acid (C8), decanoic acid (C10), hexanoic acid (C6), 10-undecenoic acid (C11Δ10), 10-bromodecanoic acid (C10Br), 10-azidodecanoic acid (C10N3). Feed volume was held at 150 mL for Trials 2-16, increased to 200 mL for Trial 17 and 250 mL for Trial 18. Trials 19-24 were reduced back to 200 mL. 4 Isopropyl β-D-1-thiogalactopyranoside (IPTG) added to Trials 7 and 8 at a final concentration of 0.5 mM to induce pBRL690 and pTrcDAEB expression, respectively.

138

4.3.6 PHA quantification

Shake flask cultures were harvested as described in Chapter 3 (Scheel et al., 2019). In brief, cells were collected by centrifugation at 3716 × g for 15 minutes, washed once with 45 mL of 35% ethanol and once with 45 mL of Nanopure filtered water, frozen at -80 °C, and dried via lyophilization.

The 10 mL culture samples from bioreactor fermentations were thawed and pelleted by centrifugation at 7,000 × g for 15 min. The supernatant was reserved and frozen at -80 °C for lyophilization and later analysis. Cell pellets were washed once with 40% ethanol (40 mL) and once with Nanopure filtered water (40 mL), pelleting with the same centrifugation conditions after each wash. Pellets were then resuspended in 5 mL Nanopure filtered water, frozen at -80 °C, and dried via lyophilization.

The PHA yields and repeating unit compositions were analyzed by GC as described in

Chapter 3 (Scheel et al., 2019). In brief, lyophilized cells (10-15 mg) were suspended in 2 mL of a 15% (v/v) sulfuric acid solution in methanol and 2 mL of chloroform and heated at 100 °C for

140 min in a 10 mL pressure vial (Kimax). The samples were cooled to room temperature, and 1 mL of Nanopure filtered water and 500 μL of methyl octanoate standard (0.25% v/v) in chloroform were added and mixed by vortex. Aqueous and organic layers were separated by centrifugation for

5 minutes at 700 rpm (Marathon 6K, Fisher Scientific). The organic layer was passed through a

0.2 μm polytetrafluoroethylene (PTFE) filter using a vacuum manifold (Millex Samplicity®) into

2 mL GC vials. Samples were injected and separated using a GC 2010 Gas Chromatograph

(Shimadzu) with an AOC-20i autoinjector and a flame ionization detector. Shimadzu’s

GCSolution software was used to analyze the data based on 3-hydroxyacyl methyl ester standard curves (except for 3-hydroxy-10-azidodecanoate monomers, which is described below).

139

4.3.7 10-azidodecanoic acid synthesis

The 10-azidodecanoic acid substrate was synthesized following previously described methods (Riva et al., 2014), and is summarized in Scheme 4.3.1. Sodium azide (15.01 g, 231 mmol) was added to a solution of 10-bromodecanoic acid (1, 40.00 g, 159 mmol) in dimethyl sulfoxide (1000 mL). The mixture was stirred at room temperature for 24 hours and then water

(2000 mL) was added. The aqueous solution was extracted with ethyl acetate (4 x 700 mL) and then washed with half-saturated brine (6 x 500mL) and brine (2 x 400mL) and dried over sodium sulfate. The sodium sulfate was filtered off and ethyl acetate was removed in vacuo to yield the azide (2, 32.84 g, 94%).

Scheme 4.3.1: Synthesis of 10-azidodecanoic acid.

4.3.8 PHA purification and molecular weight analysis

Bioreactor cultures were harvested after 48 h by dividing the volume equally among 800- mL centrifuge buckets and centrifuging at 3,716 × g for 45 min. Cell pellets were washed once with 40% ethanol (approximately 400 mL) and once with Nanopure filtered water (400 mL), pelleting with the same centrifugation conditions after each wash. Final cell pellets were transferred to a tared 250-mL centrifuge tube, frozen at -80 °C, and dried via lyophilization.

PHA was isolated from dried cells by Soxhlet extraction. Approximately 5-10 g of cells were transferred to 80 × 25 mm cellulose extraction thimbles (Whatman) and refluxed with 150 mL chloroform for 6 h. Crude polymer was concentrated by rotary evaporation and purified by

140 dropwise addition to stirring methanol (180 mL, 4°C) without exceeding a 1:10 ratio of chloroform to methanol. The methanol solution was centrifuged at 7,000 × g and 4°C for 20 min to collect any suspended polymer. This methanol purification process was repeated twice, and purified PHA was dissolved in chloroform and passed through a syringe filter (0.45 μm PTFE) before drying in vacuo.

The weight average (Mw) and number average (Mn) molecular weights for each sample were determined by gel permeation chromatography (GPC) as described in Chapter 2 (Scheel et al., 2016). In brief, PHA solutions of approximately 5.0 g L-1 were prepared by dissolution in chloroform and passed through a syringe filter (0.45 μm PTFE). Samples were injected (50 μL) into a Shimadzu LC-20AD liquid chromatograph equipped with a Shimadzu SIL-20A autosampler, a Shimadzu CTO-20A column oven, and a Shimadzu RID-10A refractive index detector. Samples were passed through an 8 x 50 mm styrenedivinylbenzene (SDV) guard column

(5 μm particles; Polymer Standards Service) and an 8 x 300 mm SDV analytical column (5 μm particles; mixed bed porosity; max molecular weight 1E6 Da; Polymer Standards Service product sda0830051lim). The column oven was maintained at 40°C with a 1 mL min-1 mobile phase of chloroform. Molecular weight standards of polystyrene with a narrow polydispersity index were used for calibration. LCsolution software (Shimadzu) was used for data analysis.

The presence of functional groups on applicable PHAs was verified by 1H-NMR spectroscopy using either a Bruker AVANCE III 600 MHz or Bruker AVANCE III HD 800 MHz instrument as noted. 1H-NMR was also used to confirm PHA production from the first successful bioreactor fermentation (Trial 9, Table 4.2). Spectra were processed with Bruker TopSpin v3.5pI2.

141

4.3.9 Calculations

PHA yields (g L-1) were calculated from the weight percent obtained using GC calibration curves as described above and the dried cell weight of 10.0 mL samples, dividing by the sample volume:

푊푒𝑖푔ℎ푡 % × 퐷푟𝑖푒푑 퐶푒푙푙 푊푒𝑖푔ℎ푡 (푔) 푃퐻퐴 푦𝑖푒푙푑 (푔 퐿−1) = ( ) 0.010 퐿

The total PHA accumulated for each bioreactor fermentation was calculated by multiplying the

PHA yield by the final culture volume from the bioreactor:

푇표푡푎푙 푃퐻퐴 (푔) = 푃퐻퐴 푦𝑖푒푙푑 (푔 퐿−1) × 퐹𝑖푛푎푙 푣표푙푢푚푒 (퐿)

Maximum theoretical yields of PHA were calculated from the total mass of fatty acid substrate available, multiplied by the molecular weight ratio of diradical monomeric equivalents to fatty acid precursor, assuming complete 1:1 conversion:

−1 푀푊푑푖푟푎푑푖푐푎푙 푚표푛표푚푒푟 (푔 푚표푙 ) 푇ℎ푒표푟푒푡𝑖푐푎푙 (푔) = 퐹푎푡푡푦 푎푐𝑖푑 푠푢푏푠푡푟푎푡푒 (푔) × ( −1 ) 푀푊푓푎푡푡푦 푎푐푖푑 (푔 푚표푙 )

The mole fraction (χ) of 3-hydroxy-10-azidodecanoate (DN3) monomers in the PHODN3 copolymer was calculated via 1H-NMR (Appendix A). Peak integrations (A), calibrated to the three protons of the terminal methyl group on 3-hydroxyoctanoate (3HO) monomers (δ 0.88), were used to determine the ratio of the DN3 protons in the alpha position (with respect to the azido group, δ 3.2, 2H) to the sum of the integrations of both chemical shifts (Sperling, 2015):

퐴3.2 2 휒퐷푁3 = 퐴3.2 퐴0.88 2 + 2

142

-1 The PHODN3 yield (g L ) for trial 24 (Table 4.2) was calculated by first obtaining the yield of the 3HO fraction using GC as described above. This was multiplied by the ratio of the mole fractions of DN3 and 3HO to find the yield of the DN3 fraction, and added to the yield of 3HO:

(푀푊 × 휒 ) −1 −1 퐷푁3 퐷푁3 −1 푃퐻푂퐷푁3 푦𝑖푒푙푑 (푔 퐿 ) = 3퐻푂 (푔 퐿 ) × + 3퐻푂 (푔 퐿 ) (푀푊3퐻푂 × 휒3퐻푂)

4.4 Results

4.4.1 Preliminary trials and the effect of pH

The goal of this investigation was to enhance the production of MCL PHA by developing high density fermentation methods for our engineered production platform, E. coli LSBJ. Previous attempts by this lab have been unsuccessful using LSBJ (unpublished data). Preliminary trials

(trials 1-8, Table 4.2) sought to address this by generating several further mutations hypothesized to eliminate negative regulation of the PHA biosynthesis pathway. The mutant strains LSBJ CRP*,

SM02, SM04, and SM06 were developed and cultured by high-density fermentation in an attempt to produce either poly(3-hydroxyoctanoate), poly(3-hydroxydecanoate), or poly(lactate-co-3- hydroxyoctanoate) (PHO, PHD, PLA-co-PHO; Fig. 4.2A). Despite some improvement in biomass accumulation, no observable polymer was produced in any of these experiments.

Based on previous observations that LSBJ would not produce PHA during bioreactor fermentation unless the pH was left uncontrolled, LSBJ was cultured in baffled shake flasks supplemented with decanoic acid using typical PHA biosynthesis conditions (Chapter 2), and both pH and PHA yield were measured every 12 h during growth (Fig. 4.2B). The pH of culture media increased substantially over 48 h to a final pH of 8.8. PHA was present after 12 h at 6.01% of dry cell weight, and more than tripled to 19.7% by 24 h while the pH increased to 8.5. The amount of

143

PHA continued to increase at a pH between 8.5 and 8.8 for the remainder of the experiment. These results suggest a correlation between an alkaline pH and PHA biosynthesis from fatty acids.

Based on our findings from the shake flask experiment, LSBJ harboring pBBRC1J4SII was cultured in the bioreactor and supplemented with decanoic acid, and upon initiating Phase 2 the pH was increased to 8.0 and maintained there for the duration of the fermentation. PHD was observed by GC analysis at a concentration of 0.16 g L-1, which was confirmed by NMR spectroscopy on purified polymers obtained by Soxhlet extraction due to the presence of characteristic chemical shifts at 5.15 and 2.5 ppm (Fig. 4.2C).

144

Figure 4.2: Summary of initial bioreactor experiments leading to successful PHD biosynthesis. (A) Results of preliminary fermentations that did not result in polymer accumulation. (B) Culture pH and PHA accumulation in LSBJ harboring pBBRC1J4SII over 48 h growth in shake flasks, sampling every 12 h. PHA was measured as a percent of dry cell weight. Data shown is the average of triplicate cultures; errors bars represent minus 1 standard deviation. (C) Photograph and 1H NMR (600 MHz) spectrum of PHD extracted from LSBJ-pBBRC1J4SII bioreactor fermentation (Trial 9). Diagnostic chemical shift for PHA at δ 5.17 – 5.13 ppm (pentet, 1H); 2.59 – 2.45 (multiplet, 2H), 1.61 – 1.47 (multiplet, 2H), 1.33 – 1.18 (multiplet, 10H), 0.8 – 0.9 (triplet, 3H). Integration indicative of residual fatty acid.

145

4.4.2 Optimization of PHD production

After successfully producing PHD in trial 9, mutant strains RSC02 (ΔarcA) and LSBJ

CRP* were analyzed to determine the effects of removing these transcriptional regulators (Trials

10 & 11, Table 4.2). Little difference was observed between LSBJ, RSC02, and LSBJ CRP*; 0.48 g L-1, 0.39 g L-1, and 0.44 g L-1 of PHD were produced by each strain after 48 h, respectively, with no observable difference between 24 h and 48 h, and high variation between duplicate fermentations (Fig. 4.3B). Dry biomass yield decreased for both mutants relative to LSBJ, reflecting a higher weight percent of PHD, and biomass decreased between 24 h and 48 h for all three strains (Fig. 4.3A). Total polymer accumulation after 48 h was low relative to the maximum theoretical yield, ranging from 5.6 to 8.3% with high variation between duplicate fermentations

(Fig. 4.3C).

Next, we reduced the rate of glucose feed addition during Phase 2 to 9.7 mL h-1 and cultured

LSBJ harboring pBBRC1J4SII to determine the effect on PHD biosynthesis (LSBJ-2, Fig. 4.3).

Biomass yield increased by approximately 50% at 24 h relative to LSBJ (29.6 g L-1); biomass still decreased between 24 h and 48 h to a final concentration of 21.3 g L-1 (Fig. 4.3A). PHD accumulation increased considerably to 2.48 g L-1 after 48 h, with no observable difference between 24 h and 48 h (Fig. 4.3B). Total polymer accumulation after 48 h was 2.60 g, a yield of

43.8 % (Fig. 4.3C).

With bioreactor fermentations continuing to show a decrease of biomass and unchanged

PHA accumulation between 24 h and 48 h, believed to be due to culture death, the bioreactor medium was changed to that described in high-density fed-batch production to support greater growth of LSBJ harboring pBBRC1J4SII (LSBJ-3, Fig. 4.3). Biomass yield increased relative to

LSBJ-2 to 41.3 and 52.0 g L-1 at 24 and 48 h, respectively (Fig. 4.3A). PHD accumulated at a

146 concentration of 2.27 g L-1 by 24 h, which doubled to 4.54 g L-1 by 48 h (Fig. 4.3B). Total polymer accumulation after 48 h was 5.12 g, a yield of 87.5% (Fig. 4.3C).

With PHD yields approaching the theoretical maximum, we increased the amount of decanoic acid added to the feed solution from 6.0 to 12.0 g (LSBJ-3 (12g), Fig. 4.3). This resulted in increased PHD yields of 3.87 and 6.18 g L-1 at 24 and 48 h, respectively, with a total yield of

7.35 g, or 62.0% of the theoretical maximum (Fig. 4.3BC). Biomass accumulation decreased slightly relative to LSBJ-2 in response to the additional fatty acid (Fig. 4.3A). To further support cell growth and PHA biosynthesis, the amount of yeast extract added to the culture medium was increased from 1.5 to 3.0 g (LSBJ-4, Fig. 4.3). This again resulted in increased PHD yields, up to

5.38 and 10.1 g L-1 at 24 h and 48 h, bringing the total yield to 10.6 g (89.7%). As the amount of decanoic acid was increased from 12 g to 25 g over subsequent trials, PHD accumulation continued to increase as well until reaching 20.1 g L-1 at 48 h, a total yield of 22.1 g (89.4%), at which point excessive foaming became an issue and the optimization process was halted (Fig. 4.3BC).

Purified polymers from each optimization experiment were analyzed by GPC to determine the number average molecular weight (Mn), weight average molecular weight (Mw), and polydispersity index (PDI, Mw/Mn) (Table 4.3). Mn values ranged from 28.6 to 64.3 kDa with no discernable trend and a high degree of variation. The polydispersity index was similarly variable, ranging from 2.3 to 3.9.

147

Figure 4.3: Optimization of PHD biosynthesis. (A) Biomass) and (B) PHD accumulation (g L1) at 24 (light grey) and 48 h (dark grey) from bioreactor fermentations. (C) Total PHD yield in grams (dark grey), displayed as an overlay of the maximum theoretical yield (light grey). All data shown is the mean of duplicate trials, plus 1 standard deviation about that mean, except for where n = 1.

148

Table 4.3: PHD Molecular Weight from Optimization Experiments.

PDI Experiment Mn (kDa) Mw (kDa) (Mw/Mn)

LSBJ 28.6 ± 13.9 81.5 ± 54.1 2.7 ± 0.6

RSC02 64.3 ± 25.4 157 ± 23.7 2.6 ± 0.6

LSBJ CRP* 40.2 ± 6.4 127 ± 4.4 3.2 ± 0.4

LSBJ-2 31.6 ± 9.8 80.4 ± 28.7 2.5 ± 0.1

LSBJ-3 35.9 ± 4.8 81.0 ± 9.5 2.3 ± 0.1

LSBJ-4 (12g) 44.7 ± 4.7 125 ± 4.4 2.8 ± 0.4

LSBJ-4 (15g) 37.3 ± 0.1 117 ± 0.6 3.1 ± 0

LSBJ-4 (20g) 37.3 ± 4.8 103 ± 26.0 2.8 ± 0.4

LSBJ-4 (25g) 42.0 ± 3.7 164 ± 43.7 3.9 ± 0.7

4.4.3 Production of alternative MCL PHA

To further investigate the capabilities of the optimized high-density fermentation process,

LSBJ harboring pBBRC1J4SII was supplemented with two other saturated MCL fatty acids in place of decanoic acid: hexanoic acid and octanoic acid (Trials 19 & 20, Table 4.2). Feed rates and fatty acid concentrations (20 g in 200 mL) were kept the same as Trial 17 to avoid the foaming issue observed at higher decanoic acid concentrations (Trial 18). High biomass accumulation was observed for both trials, similar to biomass yields obtained for Trial 17 (Fig. 4.4A). Both fatty acids were successfully converted into their corresponding polymers, PHHx and PHO, with yields of 18.2 and 17.3 g L-1 after 48 h, respectively, and followed a similar trend of increased PHA accumulation between 24 h and 48 h (Fig. 4.4B). Total polymer accumulation approached the maximum theoretical yield for both PHHx and PHO (99.0% and 90.8%) (Fig. 4.4C). The

149 molecular weights of the resulting polymers were higher than those obtained for PHD, while the

PDI remained consistent (Table 4.4).

Biosynthesis of functionalized PHAs, desirable materials that are more challenging to produce, was also attempted using the optimized high-density fermentation process (Trial 21 &

22, Table 4.2). Utilizing the same culture conditions and inoculum, 10-undecenoic acid and 10- bromodecanoic acid were supplemented to produce their corresponding polymers, poly(3- hydroxy-10-undecenoate) (PHU) and poly(3-hydroxy-10-bromodecanoate) (PHDBr) (Fig. 4.4).

PHU yield was lower at 48 h compared to other polymers at 9.24 g L-1, although due to excessive foaming and subsequent antifoam addition the final culture volume was high and the total polymer accumulation was 13.2 g (66.8% of theoretical, Fig. 4.4BC). Similarly, 10-bromodecanoic acid caused excessive foaming and resulted in a lower yield and higher final volume, although to a lesser extent; PHDBr yield at 48 h was 14.3 g L-1 with a total accumulation of 18.0 g (90.7%). The

Mn of PHU was similar to that of PHD at 39.1 kDa, while the Mn of PHDBr was low at 21.6 kDa

(Table 4.4).

Biosynthesis of copolymers was also demonstrated using the high-density fermentation process by supplementing the feed with two fatty acids simultaneously. Octanoic acid was first mixed with decanoic acid (Trial 23), then with 10-azidodecanoic acid (Trial 24), both with 80 mol% octanoic acid (0.58 M total fatty acid). The copolymers poly(3-hydroxyoctanoate-co-3- hydroxydecanoate) (PHOD) and poly(3-hydroxyoctanoate-co-3-hydroxy-10-azidodecanoate)

-1 -1 (PHODN3) were both produced successfully, with yields at 48 h of 16.7 g L and 18.7 g L and total accumulation of 16.7 (97.4%) and 17.8 g (97.9%), respectively (Fig. 4.4BC). The Mn of

PHOD was 105 kDa, and the Mn of PHODN3 was 47.1 kDa (Table 4.4).

150

Figure 4.4: Alternative MCL PHA biosynthesis. (A) Biomass) and (B) PHD accumulation (g L1) at 24 (light grey) and 48 h (dark grey) from bioreactor fermentations. (C) Total PHD yield in grams (dark grey), displayed as an overlay of the maximum theoretical yield (light grey). Trials were not performed with replicates and demonstrate proof-of-concept for expanding the possible substrates for the optimized high-density fermentation process.

151

Table 4.4: Alternative MCL PHA Molecular Weights

Polymer Mn (kDa) Mw (kDa) PDI

PHHx 117.8 334.9 2.8

PHO 66.3 192.3 2.9

PHU 39.1 132.9 3.4

PHDBr 21.6 70.5 3.3

PHOD 104.7 286.7 2.7

PHODN3 47.1 149.0 3.1

4.5 Discussion

The goal of this study was to develop a fed-batch high-density fermentation process capable of producing MCL PHA with controlled monomer compositions at a higher titer than that previously achieved in shake flasks. Based on the evidence in Fig. 4.3 and Fig. 4.4b, this process offers a significant improvement over previous shake flask studies, and a variety of MCL PHA homopolymers and copolymers were made successfully (Scheel et al., 2016; Tappel et al., 2012b).

In initial experiments with high-density fermentation of LSBJ, including those published here as well as unpublished data from our research group, we were unable to detect any evidence of PHA biosynthesis. Preliminary experiments sought to address this through a series of genetic engineering experiments, beginning with the creation of LSBJ CRP*. In LSBJ, which has a wild- type CRP, catabolite repression plays an important role in regulating primary and secondary metabolism (Franchini et al., 2015; Perrenoud and Sauer, 2005). In the bioreactor fermentation, which used glucose as the carbon source, we hypothesized that low levels of CRP-cAMP were preventing the expression of β-oxidation genes necessary for PHA biosynthesis, and that

152 conferring the CRP* mutation to LSBJ would alleviate this repression (Clark, 1981; Khankal et al., 2009; Wade et al., 2001). This was not found to be the case in initial experiments (data not shown), and further mutations were made to this strain in the subsequent preliminary experiments

(Trials 1-8, Table 4.2, Fig. 4.2A).

Production of PLA-co-PHO, a novel copolymer that has generated research interest, was attempted in Trial 1 in conjunction with an ongoing project in the Nomura Research Group.

Deletion of the dld gene from LSBJ CRP*, which encodes for D-lactate dehydrogenase and has been shown to prevent the accumulation of PLA in E. coli (Taguchi et al., 2008), did not result in any detectable PLA-co-PHO. Although attempts to incorporate lactate monomers were abandoned in subsequent trials, the Δdld mutant (SM02) was further engineered by deletion of the arcA gene.

Deletion of arcA was previously shown to upregulate expression of β-oxidation genes and enhance

MCL PHA biosynthesis, and we hypothesized that the removal of arcA combined with the CRP* mutation (SM04) would enable PHA biosynthesis during bioreactor fermentation (Scheel et al.,

2016). Again, this was not supported by the evidence, with no detectable polymer observed (Trial

2).

We next hypothesized that the exogenous octanoic acid was being directed towards fatty acid biosynthesis rather than degradation, and that this could be prevented by deletion of the aas gene encoding for acyl-acyl carrier protein (acyl-ACP) synthetase. Acyl-ACP synthetase has a high affinity towards octanoate, and coupled with the increased membrane biosynthesis necessary for rapid culture growth during high-density fermentation it is possible that the octanoic acid was routed away from β-oxidation (Black and DiRusso, 2003; Ray and Cronan, 1976). However, the aas-deficient strain SM06 was unable to produce PHO from octanoic acid (Trial 3). Increasing the amount of octanoic acid available, both in bulk addition and continuous feeding strategies with

153

SM06, likewise did not result in polymer accumulation (Trial 4-6). A stronger promoter, generated by converting the lac promoter of pBBRC1J4SII into the tac promoter by mutagenesis of the -35 region, was hypothesized to enable PHA biosynthesis by increasing expression of phaJ4 and phaC1 (Trial 7). This did not work, however this is unsurprising as those two genes are under the control of a low-expressing constitutive Cupriavidus necator promoter inserted before the lac promoter of pBBRC1J4SII, and PHA biosynthesis has been demonstrated with this system numerous times during growth in shake flasks (Tappel et al., 2012b; Wang et al., 2012).

Overexpression of the atoDAEB operon, which was previously shown to occur in the arcA- deficient LSBJ strain RSC02 (Scheel et al., 2016), was also unsuccessful in enabling PHA biosynthesis (Trial 8).

The effect of increasing the culture pH to 8.0 during Phase 2 is simultaneously the most profound and confounding result from this study (Trial 9). It was essential to the production of

MCL PHA from exogenous fatty acids in LSBJ (Fig. 4.2C); maintaining the pH at 7.0 prevented

PHA accumulation in all preliminary trials (Fig. 4.2A), as well as numerous trials not shown here.

It is well-documented that growing E. coli to stationary phase in LB (and other tryptone-based media) commonly results in alkalinization, which explains the results of Fig. 4.2B and could explain why shake flask production is successful at producing MCL PHA. The literature does not offer any clear connections between MCL fatty acid uptake and catabolism at an alkaline pH, although there is a paucity of research in this area relative to growth conditions at a neutral pH

(Maurer et al., 2005; Padan et al., 2005; Stancik et al., 2002; Yohannes et al., 2004). Despite no clear connections, several aspects of E. coli pH homeostasis in alkaline environments may have an indirect effect on PHA biosynthesis. Numerous transport systems are up-regulated to direct the

H+ flow into the cell, and FadD does transport H+ concomitantly with fatty acids (Weimar et al.,

154

2002). No evidence of differential expression for fadD has been observed, however transcriptome studies can not capture proteins that are regulated at the activity level rather than expression, which is common for many Na+/H+ antiporters (Maurer et al., 2005; Padan et al., 2005). Another contributing factor could be the repression of the flagellar biosynthesis and chemotaxis regulons, which are down-regulated extensively (47 out of 50 genes, pH 8.7) at an alkaline pH (Maurer et al., 2005). Repression of this energy-intensive process could have freed up cellular resources such as ATP for PHA biosynthesis. Further research into fatty acid uptake and metabolism would be necessary to determine the direct effects of pH on PHA biosynthesis, which falls far beyond the scope of this study.

Although successful, the PHD yield in Trial 9 did not offer an improvement over typical shake flask production. To test the earlier hypotheses of low CRP-cAMP inhibiting biosynthesis and repression by ArcA, we repeated the conditions used in Trial 9 with both RSC02 (Trial 10) and LSBJ CRP* (Trial 11). No benefit to PHD yield was observed for either Trial; PHD weight percent was improved slightly, but this was offset by a reduced biomass (Fig. 4.3AB).

Reducing the rate of glucose addition during Phase 2 (Trial 12) increased the yield of PHD to 2.48 g L-1, greater than a 5-fold improvement (LSBJ-2, Fig. 4.3B). One explanation for this is the overabundance of glucose in previous trials and the subsequent increase in overflow metabolism. Overflow metabolism in E. coli occurs during high levels of aerobic glucose consumption, and results in the excretion of acetate which can become toxic and growth-limiting

(Vemuri et al., 2006). This is particularly pronounced in E. coli K-12 strains, which LSBJ is derived from, due to an inactive or low-activity glyoxylate shunt (Noronha et al., 2000; Rand et al., 2017; Walle and Shiloach, 1998). An increase in biomass of approximately 50% provides additional evidence to support this explanation, however this biomass increase is not enough to

155 account for the magnitude of the increase in PHD yield (Fig. 4.3A). Another effect of the reduced glucose feed rate could be an increase in the cAMP pool resulting from increased adenylate cyclase activity due to decreased glucose transport (Notley-McRobb et al., 1997). This in turn would increase CRP-cAMP abundance and could lead to upregulation of fadD and fadE, resulting in enhanced PHD biosynthesis.

One issue that needed to be addressed was the apparent death of the bacterial culture during

Phase 2, as evidenced by a reduction in biomass and lack of PHA accumulation between 24 and

48 h (Fig. 4.3AB), and by a marked decrease in oxygen consumption as early as 13 h until DO% remained nearly saturated (Fig. D1-6, Appendix D). Previous research has indicated that high ammonia concentrations during high-density fermentation inhibit the growth of E. coli and limit secondary metabolism, which could explain the results observed in this study (Lau et al., 2004;

Riesenberg et al., 1991; Shiloach and Fass, 2005). Reformulating the culture medium and glucose feed based on low ammonia alternatives described by Lau et al. (2004) appeared to eliminate the issue of culture death (Trial 13); biomass and PHD yield increased between 24 and 48 h (Fig.

4.3AB), oxygen demand slowly increased over the entire fermentation (Fig. D7, Appendix D), and total PHD accumulation at 48 h reached 87.5% of the theoretical yield (Fig. 4.3C). In addition to drastically reducing the ammonium concentration, the new formulation contained more phosphate to support a higher cell density and more trace metals to ensure adequate cofactor abundance

(Herbert et al., 1965; Neidhardt et al., 1974).

With a media formulation capable of supporting high-density fermentation and the majority of fatty acid substrate being converted to polymer, the next logical step was the optimization of the fatty acid feed; specifically, increasing the amount and rate of decanoic acid addition (Trials

14-18). Progressively increasing the amount of decanoic acid added to totals of 12 g, 15 g, 20 g,

156 and 25 g (addition rates of 0.44, 0.55, 0.65, and 0.75 g h-1, respectively) correlated with an increase in PHD yield, culminating in a final titer of 20.1 g L-1 (Fig. 4.3B). However, adding 25 g of decanoic acid resulted in excessive foaming likely as a result of exogenous decanoate behaving as a surfactant, and replication of this feeding strategy proved to be difficult due to uncontrollable foaming and loss of culture. Alternative strategies exist to control foam in an industrial bioprocess setting, such as mechanical disruption or ultrasound, but these are incompatible with the small benchtop vessel used in this study (Delvigne and Lecomte, 2009; Routledge, 2012). As such the optimization of the fatty acid feed was halted, and analysis of previous metabolic engineering strategies (CRP*, ΔarcA, etc.) was deemed unnecessary due to the physical nature of the foaming limitation. It is also important to note that the amount of yeast extract added to the initial culture medium was doubled to 3.0 g for Trial 15 and for each subsequent Trial. This was in response to a decrease in % yield when decanoic acid was increased to 12 g, believed to be due to an increased fatty acid toxicity. Fadzil et al. has previously shown that increasing the concentration of yeast extract for E. coli LSBJ cultures utilizing co-carbon sources (glucose and decanoic acid) increases their tolerance towards fatty acids and enhances MCL PHA biosynthesis (2018). Our results support this observation; increasing the concentration of yeast extract led to improved biomass,

PHD yield, and decanoic acid utilization (89.7% yield) (LSBJ-4 (12g), Fig. 4.3).

The optimized high-density fermentation process developed here was also shown to be flexible with regard to fatty acid substrate; hexanoic acid, octanoic acid, 10-undecenoic acid, 10- bromodecanoic acid, and 10-azidodecanoic acid were all successfully polymerized in addition to decanoic acid (Fig. 4.4), with confirmation by NMR for the functionalized PHU, PHDBr, and

-1 PHODN3 (Appendix A). Each alternative MCL PHA tested was produced at yields of 9.24 g L or higher and with no detectable co-monomers present except for the intentional copolymers

157

PHOD and PHODN3. These copolymers contained monomer compositions closely resembling the mol% of the feed solution (81.8 and 80.4% 3HO, respectively), which is unsurprising given the high fatty acid utilization observed (approximately 97% for both) (Fig. 4.4C). This control over monomer composition is one of the primary benefits to using E. coli LSBJ as the production host, particularly over a wide range of fatty acid substrates (Tappel et al., 2012b). Although numerous high-density cultivation methods have been reported using native producers such as Pseudomonas species, only PHD has previously been produced at near-homopolymeric compositions and high titers, and there are no reported bioprocesses for the production of PHHx, PHO, PHU, or PHDBr homopolymers to the best of our knowledge (Follonier et al., 2015; Gao et al., 2018; Jiang et al.,

2013; Oliveira et al., 2020; Sun et al., 2007).

The molecular weights varied for the PHAs produced, and the PDIs ranged from 2.3 to 3.9

(Table 4.3 & 4.4). The molecular weights (Mn) of PHD showed considerable variation, ranging between 28.6 and 64.3 kDa. This is comparable and slightly less than previous work performed in shake flasks, where Mn values of 61.5 to 111 kDa were observed (Mohd Fadzil et al., 2018; Tappel et al., 2012b). Similarly, the Mn value of PHU was slightly less than previously observed (39.1 vs.

60 kDa) (Tappel et al., 2012a), while the Mn of PHHx was higher than previous reports (118 vs.

77 kDa) (Scheel et al., 2016). Overall molecular weights are comparable to those from similar studies, however with the high degree of variation and lack of replicates for alternative MCL PHAs only limited comparisons can be made.

The high-density fermentation process described here was developed to allow for MCL

PHA biosynthesis with controlled monomer composition at a higher yield than shake flask production can accommodate. The productivity shown here of PHD biosynthesis from 25 g decanoic acid (0.42 g L-1 h-1) rivals that of the highest published productivity for nearly

158 homopolymeric PHD (0.41 g L-1 h-1), although this work had issues with reproducibility (Gao et al., 2018). High MCL PHA yields were demonstrated from a variety of fatty acid substrates, both saturated and functionalized, showing that this is a versatile process capable of producing unique and valuable materials.

4.6 References

Black, P.N., DiRusso, C.C., 2003. Transmembrane Movement of Exogenous Long-Chain Fatty

Acids: Proteins, Enzymes, and Vectorial Esterification. Microbiol. Mol. Biol. Rev. 67,

454–472. https://doi.org/10.1128/MMBR.67.3.454-472.2003

Blunt, W., Levin, D.B., Cicek, N., 2018. Bioreactor Operating Strategies for Improved

Polyhydroxyalkanoate (PHA) Productivity. Polymers 10, 1197.

https://doi.org/10.3390/polym10111197

Clark, D., 1981. Regulation of fatty acid degradation in Escherichia coli: analysis by operon

fusion. J. Bacteriol. 148, 521–526.

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

coli K-12 using PCR products. Proceedings of the National Academy of Sciences 97,

6640–6645.

Delvigne, F., Lecomte, J., 2009. "Foam formation and control in bioreactors" Encyclopedia of

Industrial Biotechnology. John Wiley & Sons, Inc.

Deutscher, J., 2008. The mechanisms of carbon catabolite repression in bacteria. Curr. Opin.

Microbiol. 11, 87–93. https://doi.org/10.1016/j.mib.2008.02.007

Feng, Y., Cronan, J.E., 2012. Crosstalk of Escherichia coli FadR with Global Regulators in

Expression of Fatty Acid Transport Genes. PLoS ONE 7, e46275.

https://doi.org/10.1371/journal.pone.0046275

159

Fic, E., Bonarek, P., Gorecki, A., Kedracka-Krok, S., Mikolajczak, J., Polit, A., Tworzydlo, M.,

Dziedzicka-Wasylewska, M., Wasylewski, Z., 2009. cAMP Receptor Protein from

Escherichia coli as a Model of Signal Transduction in Proteins - A Review. Journal of

Molecular Microbiology and Biotechnology 17, 1–11. https://doi.org/10.1159/000178014

Follonier, S., Riesen, R., Zinn, M., 2015. Pilot-scale Production of Functionalized mcl-PHA from

Grape Pomace Supplemented with Fatty Acids. Chemical and Biochemical Engineering

Quarterly 29, 113–121. https://doi.org/10.15255/CABEQ.2014.2251

Franchini, A.G., Ihssen, J., Egli, T., 2015. Effect of Global Regulators RpoS and Cyclic-AMP/CRP

on the Catabolome and Transcriptome of Escherichia coli K12 during Carbon- and Energy-

Limited Growth. PLoS ONE 10, e0133793. https://doi.org/10.1371/journal.pone.0133793

Gao, J., Vo, M.T., Ramsay, J.A., Ramsay, B.A., 2018. Overproduction of MCL-PHA with high 3-

hydroxydecanoate Content. Biotechnology and Bioengineering 115, 390–400.

https://doi.org/10.1002/bit.26474

Guidi-Rontani, C., Danchin, A., Ullmann, A., 1981. Isolation and characterization of an

Escherichia coli mutant affected in the regulation of adenylate cyclase. J. Bacteriol. 148,

753–761.

Herbert, D., Phipps, P.J., Tempest, D.W., 1965. The chemostat: design and instrumentation. Lab

Pract 14, 1150–1161.

Iram, S.H., Cronan, J.E., 2006. The β-Oxidation Systems of Escherichia coli and Salmonella

enterica Are Not Functionally Equivalent. Journal of Bacteriology 188, 599–608.

https://doi.org/10.1128/JB.188.2.599-608.2006

160

Isaacs, H., Chao, D., Yanofsky, C., Saier, M.H., 1994. Mechanism of catabolite repression of

tryptophanase synthesis in Escherichia coli. Microbiology 140 (8), 2125–2134.

https://doi.org/10.1099/13500872-140-8-2125

Jiang, X.J., Sun, Z., Ramsay, J.A., Ramsay, B.A., 2013. Fed-batch production of MCL-PHA with

elevated 3-hydroxynonanoate content. AMB Express 3, 50. https://doi.org/10.1186/2191-

0855-3-50

Kameda, K., Nunn, W.D., 1981. Purification and characterization of acyl coenzyme A synthetase

from Escherichia coli. J. Biol. Chem. 256, 5702–5707.

Khankal, R., Chin, J.W., Ghosh, D., Cirino, P.C., 2009. Transcriptional effects of CRP* expression

in Escherichia coli. Journal of Biological Engineering 3, 13. https://doi.org/10.1186/1754-

1611-3-13

Koller, M., 2018. A Review on Established and Emerging Fermentation Schemes for Microbial

Production of Polyhydroxyalkanoate (PHA) Biopolyesters. Fermentation 4, 30.

https://doi.org/10.3390/fermentation4020030

Lau, J., Tran, C., Licari, P., Galazzo, J., 2004. Development of a high cell-density fed-batch

bioprocess for the heterologous production of 6-deoxyerythronolide B in Escherichia coli.

J. Biotechnol. 110, 95–103. https://doi.org/10.1016/j.jbiotec.2004.02.001

Lee, S.Y., Choi, J., Wong, H.H., 1999. Recent advances in polyhydroxyalkanoate production by

bacterial fermentation: mini-review. International Journal of Biological Macromolecules

25, 31–36. https://doi.org/10.1016/S0141-8130(99)00012-4

Lepore, B.W., Indic, M., Pham, H., Hearn, E.M., Patel, D.R., van den Berg, B., 2011. Ligand-

gated diffusion across the bacterial outer membrane. Proc Natl Acad Sci U S A 108, 10121–

10126. https://doi.org/10.1073/pnas.1018532108

161

Levine, A.C., Sparano, A., Twigg, F.F., Numata, K., Nomura, C.T., 2015. Influence of Cross-

Linking on the Physical Properties and Cytotoxicity of Polyhydroxyalkanoate (PHA)

Scaffolds for Tissue Engineering. ACS Biomaterials Science & Engineering 1, 567–576.

https://doi.org/10.1021/acsbiomaterials.5b00052

Lu, J., Tappel, R.C., Nomura, C.T., 2009. Mini-Review: Biosynthesis of Poly(hydroxyalkanoates).

Polymer Reviews 49, 226 248. https://doi.org/10.1080/15583720903048243

Maurer, L.M., Yohannes, E., Bondurant, S.S., Radmacher, M., Slonczewski, J.L., 2005. pH

Regulates Genes for Flagellar Motility, Catabolism, and Oxidative Stress in Escherichia

coli K-12. Journal of Bacteriology 187, 304–319. https://doi.org/10.1128/JB.187.1.304-

319.2005

Mohd Fadzil, F.I., Mizuno, S., Hiroe, A., Nomura, C.T., Tsuge, T., 2018. Low Carbon

Concentration Feeding Improves Medium-Chain-Length Polyhydroxyalkanoate

Production in Escherichia coli Strains With Defective β-Oxidation. Front Bioeng

Biotechnol 6. https://doi.org/10.3389/fbioe.2018.00178

Neidhardt, F.C., Bloch, P.L., Smith, D.F., 1974. Culture medium for enterobacteria. J. Bacteriol.

119, 736–747.

Noronha, S.B., Yeh, H.J.C., Spande, T.F., Shiloach, J., 2000. Investigation of the TCA cycle and

the glyoxylate shunt in Escherichia coli BL21 and JM109 using 13C-NMR/MS.

Biotechnology and Bioengineering 68, 316–327. https://doi.org/10.1002/(SICI)1097-

0290(20000505)68:3<316::AID-BIT10>3.0.CO;2-2

Notley-McRobb, L., Death, A., Ferenci, T., 1997. The relationship between external glucose

concentration and cAMP levels inside Escherichia coli: implications for models of

162

phosphotransferase-mediated regulation of adenylate cyclase. Microbiology (Reading,

Engl.) 143 ( Pt 6), 1909–1918. https://doi.org/10.1099/00221287-143-6-1909

Oliveira, G.H.D., Zaiat, M., Rodrigues, J.A.D., Ramsay, J.A., Ramsay, B.A., 2020. Towards the

Production of mcl-PHA with Enriched Dominant Monomer Content: Process Development

for the Sugarcane Biorefinery Context. J Polym Environ 28, 844–853.

https://doi.org/10.1007/s10924-019-01637-2

Padan, E., Bibi, E., Ito, M., Krulwich, T.A., 2005. Alkaline pH homeostasis in bacteria: New

insights. Biochimica et Biophysica Acta (BBA) - Biomembranes 1717, 67–88.

https://doi.org/10.1016/j.bbamem.2005.09.010

Perrenoud, A., Sauer, U., 2005. Impact of Global Transcriptional Regulation by ArcA, ArcB, Cra,

Crp, Cya, Fnr, and Mlc on Glucose Catabolism in Escherichia coli. Journal of Bacteriology

187, 3171–3179. https://doi.org/10.1128/JB.187.9.3171-3179.2005

Pfeifer, B., Hu, Z., Licari, P., Khosla, C., 2002. Process and Metabolic Strategies for Improved

Production of Escherichia coli-Derived 6-Deoxyerythronolide B. Applied and

Environmental Microbiology 68, 3287–3292. https://doi.org/10.1128/AEM.68.7.3287-

3292.2002

Pinto, A., Ciesla, J.H., Palucci, A., Sutliff, B.P., Nomura, C.T., 2016. Chemically Intractable No

More: In Vivo Incorporation of “Click”-Ready Fatty Acids into Poly-[(R)-3-

hydroxyalkanoates] in Escherichia coli. ACS Macro Letters 5, 215–219.

https://doi.org/10.1021/acsmacrolett.5b00823

Rai, R., Keshavarz, T., Roether, J.A., Boccaccini, A.R., Roy, I., 2011. Medium chain length

polyhydroxyalkanoates, promising new biomedical materials for the future. Materials

163

Science and Engineering: R: Reports 72, 29–47.

https://doi.org/10.1016/j.mser.2010.11.002

Rand, J.M., Gordon, G.C., Mehrer, C.R., Pfleger, B.F., 2017. Genome sequence and analysis of

Escherichia coli production strain LS5218. Metab Eng Commun 5, 78–83.

https://doi.org/10.1016/j.meteno.2017.10.001

Ray, T.K., Cronan, J.E., 1976. Activation of long chain fatty acids with acyl carrier protein:

demonstration of a new enzyme, acyl-acyl carrier protein synthetase, in Escherichia coli.

Proc. Natl. Acad. Sci. U.S.A. 73, 4374–4378. https://doi.org/10.1073/pnas.73.12.4374

Riesenberg, D., Schulz, V., Knorre, W.A., Pohl, H.D., Korz, D., Sanders, E.A., Ross, A., Deckwer,

W.D., 1991. High cell density cultivation of Escherichia coli at controlled specific growth

rate. J. Biotechnol. 20, 17–27. https://doi.org/10.1016/0168-1656(91)90032-q

Riva, E., Wilkening, I., Gazzola, S., Li, W.M.A., Smith, L., Leadlay, P.F., Tosin, M., 2014.

Chemical probes for the functionalization of polyketide intermediates. Angew. Chem. Int.

Ed. Engl. 53, 11944–11949. https://doi.org/10.1002/anie.201407448

Routledge, S.J., 2012. Beyond de-foaming: the effects of antifoams on bioprocess productivity.

Comput Struct Biotechnol J 3. https://doi.org/10.5936/csbj.201210014

Sambrook, J.J., Russell, D.W., 2001. Molecular cloning: A laboratory manual. Cold Spring Harbor

Laboratory Press, Cold Spring Harbor, NY.

Scheel, R.A., Fusi, A.D., Min, B.C., Thomas, C.M., Ramarao, B.V., Nomura, C.T., 2019.

Increased Production of the Value-Added Biopolymers Poly(R-3-Hydroxyalkanoate) and

Poly(γ-Glutamic Acid) From Hydrolyzed Paper Recycling Waste Fines. Front. Bioeng.

Biotechnol. 7. https://doi.org/10.3389/fbioe.2019.00409

164

Scheel, R.A., Ji, L., Lundgren, B.R., Nomura, C.T., 2016. Enhancing poly(3-hydroxyalkanoate)

production in Escherichia coli by the removal of the regulatory gene arcA. AMB Express

6, 120. https://doi.org/10.1186/s13568-016-0291-z

Shiloach, J., Fass, R., 2005. Growing E. coli to high cell density—A historical perspective on

method development. Biotechnology Advances 23, 345–357.

https://doi.org/10.1016/j.biotechadv.2005.04.004

Sperling, L.H., 2015. Introduction to Physical Polymer Science, 4th ed. John Wiley & Sons.

Stancik, L.M., Stancik, D.M., Schmidt, B., Barnhart, D.M., Yoncheva, Y.N., Slonczewski, J.L.,

2002. pH-Dependent Expression of Periplasmic Proteins and Amino Acid Catabolism in

Escherichia coli. Journal of Bacteriology 184, 4246–4258.

https://doi.org/10.1128/JB.184.15.4246-4258.2002

Sun, Z., Ramsay, J.A., Guay, M., Ramsay, B., 2007. Increasing the yield of MCL-PHA from

nonanoic acid by co-feeding glucose during the PHA accumulation stage in two-stage fed-

batch fermentations of Pseudomonas putida KT2440. Journal of Biotechnology,

International Symposium on Biological Polyesters 2006 132, 280–282.

https://doi.org/10.1016/j.jbiotec.2007.02.023

Taguchi, S., Yamada, M., Matsumoto, K., Tajima, K., Satoh, Y., Munekata, M., Ohno, K., Kohda,

K., Shimamura, T., Kambe, H., Obata, S., 2008. A microbial factory for lactate-based

polyesters using a lactate-polymerizing enzyme. PNAS 105, 17323–17327.

https://doi.org/10.1073/pnas.0805653105

Tappel, R.C., Kucharski, J.M., Mastroianni, J.M., Stipanovic, A.J., Nomura, C.T., 2012a.

Biosynthesis of Poly[(R)-3-hydroxyalkanoate] Copolymers with Controlled Repeating

Unit Compositions and Physical Properties. Biomacromolecules 13, 2964.

165

Tappel, R.C., Wang, Q., Nomura, C.T., 2012b. Precise control of repeating unit composition in

biodegradable poly(3-hydroxyalkanoate) polymers synthesized by Escherichia coli.

Journal of Bioscience and Bioengineering 113, 480–486.

https://doi.org/10.1016/j.jbiosc.2011.12.004

Vemuri, G.N., Altman, E., Sangurdekar, D.P., Khodursky, A.B., Eiteman, M.A., 2006. Overflow

Metabolism in Escherichia coli during Steady-State Growth: Transcriptional Regulation

and Effect of the Redox Ratio. Appl. Environ. Microbiol. 72, 3653–3661.

https://doi.org/10.1128/AEM.72.5.3653-3661.2006

Wade, J.T., Belyaeva, T.A., Hyde, E.I., Busby, S.J., 2001. A simple mechanism for co-dependence

on two activators at an Escherichia coli promoter. EMBO J. 20, 7160–7167.

https://doi.org/10.1093/emboj/20.24.7160

Walle, M. van de, Shiloach, J., 1998. Proposed mechanism of acetate accumulation in two

recombinant Escherichia coli strains during high density fermentation. Biotechnology and

Bioengineering 57, 71–78. https://doi.org/10.1002/(SICI)1097-

0290(19980105)57:1<71::AID-BIT9>3.0.CO;2-S

Wang, Q., Tappel, R.C., Zhu, C., Nomura, C.T., 2012. Development of a New Strategy for

Production of Medium-Chain-Length Polyhydroxyalkanoates by Recombinant

Escherichia coli via Inexpensive Non-Fatty Acid Feedstocks. Applied and Environmental

Microbiology 78, 519–527. https://doi.org/10.1128/AEM.07020-11

Weimar, J.D., DiRusso, C.C., Delio, R., Black, P.N., 2002. Functional role of fatty acyl-coenzyme

A synthetase in the transmembrane movement and activation of exogenous long-chain fatty

acids. Amino acid residues within the ATP/AMP signature motif of Escherichia coli FadD

166

are required for enzyme activity and fatty acid transport. J. Biol. Chem. 277, 29369–29376.

https://doi.org/10.1074/jbc.M107022200

Yohannes, E., Barnhart, D.M., Slonczewski, J.L., 2004. pH-Dependent Catabolic Protein

Expression during Anaerobic Growth of Escherichia coli K-12. J Bacteriol 186, 192–199.

https://doi.org/10.1128/JB.186.1.192-199.2004

167

Chapter 5: Progress towards a poly(3-hydroxyoctanoate-co-3-hydroxy-10-azidodecanoate) nanoparticle drug delivery system

5.1 Abstract

This chapter describes progress towards the production of nanoparticles made from azide- containing poly(3-hydroxyalkanote) copolymers (Azido PHAs), with the purpose of developing a targeted drug-delivery system. Two azido PHAs, poly(3-hydroxybutyrate-co-3-hydroxy-10- azidodecanoate) (PHBDN3) and poly(3-hydroxyoctanoate-co-3-hydroxy-10-azidodecanoate)

(PHODN3), were biosynthesized in recombinant Escherichia coli RSC02 harboring PHA biosynthesis genes. Bacterial endotoxin was removed from isolated polymers by a combination of sodium hydroxide treatment and methanol precipitation, which were effective individually and in combination to reduce endotoxin below 0.01 endotoxin units per mg (EU mg-1). Polymer molecular weights were largely unaffected by the sodium hydroxide treatment. Nanoparticles were iteratively produced by single emulsion or nanoprecipitation techniques to determine the most effective method for generating nanoparticles <100 nm in diameter. The ideal production method was nanoprecipitation of PHODN3 dissolved in 15% ethanol/acetone solution and precipitated in

Nanopure filtered water, which yielded particles with an average diameter of 70.6 nm. With this process, paclitaxel was encapsulated by PHODN3 with an encapsulation efficiency as high as

80.8% when 1% w/w paclitaxel was administered, with a decrease in efficiency as the concentration of paclitaxel was increased. PLGA nanoparticles were prepared by the same process, wherein encapsulation efficiencies under 14.3% were achieved. Nanoparticle uptake kinetics and cytotoxicity experiments must still be performed with human cancer cells.

168

5.2 Introduction

Polymeric nanoparticles have received considerable attention in recent years as potential drug-delivery systems, particularly for use in targeted chemotherapy treatments (Hans and

Lowman, 2002; Hillaireau and Couvreur, 2009; Lee et al., 2017). Nanoparticles are generally defined as spherical, solid particles approximately 100 nm in size. They have been produced using a variety of polymers, though most commonly with poly(lactide-co-glycolide) (PLGA) due to its biocompatibility and biodegradability into common metabolites (lactate and glycolate), commercial availability, and its approval by both the U.S. Food and Drug Administration (FDA) and European Medicine Agency (EMA) for use in therapeutic devices (extensively reviewed by

Danhier et al., 2012).

Numerous nanoparticle targeting strategies have been investigated as a way to selectively administer drugs to target cells and reduce systemic toxicity, most commonly by the conjugation of targeting moieties to the exterior particle surface (Yu et al., 2012). These strategies include; antibodies such as anti-prostate stem cell antigen antibody and mutant anti-epidermal growth factor receptor antibody (Hadjipanayis et al., 2010; Ling et al., 2011), peptides and proteins like transferrin (Petrelli et al., 2010; Sarris et al., 2000), and small molecules such as riboflavin and folic acid (Foraker et al., 2003; Kennedy et al., 2003). Folic acid in particular has generated significant research interest due to its low cost and high availability, and more importantly because of the high folate-receptor expression observed in many types of cancerous cells (Bahrami et al.,

2015). Folate receptor-α (FR-α) has a high binding affinity for folate and induces uptake by endocytosis when bound, and folate-conjugated nanoparticles have been shown to be similarly transported, offering a direct route for nanoparticle internalization (Paulos et al., 2004; Zhao et al.,

2014).

169

Poly[(R)-3-hydroxyalkanoates] (PHAs) are another class of biodegradable, biocompatible polymers that are promising biomedical materials (Rai et al., 2011). These bacterially-derived polyesters have been investigated as a potential nanoparticle drug-delivery platform, with limited success; the predominant PHAs employed for nanoparticle synthesis included the short chain- length (SCL) PHA poly(3-hydroxybutyrate) and its copolymers containing 3-hydroxyvalerate, which were characterized by low drug encapsulation efficiency, large particle size (>160 nm), and poor drug retention (Poletto et al., 2008; Pramual et al., 2016; Xiong et al., 2010). These SCL

PHAs also suffer from chemical intractability, making them difficult to functionalize with targeting moieties. Additionally, PHAs biosynthesized via gram-negative bacteria contain lipopolysaccharides (LPS, endotoxin), a component of the outer membrane of gram-negative bacteria that can induce a severe inflammatory response in humans known as endotoxic shock

(Calvano and Coyle, 2012; Rietschel et al., 1994). Endotoxins are persistent, requiring additional purification steps to remove them from PHAs (Furrer et al., 2007; Lee et al., 1999).

Recently, a number of medium chain length (MCL) PHAs have been biosynthesized containing terminal alkene, alkyne, azide, and bromo functional groups (Nkrumah-Agyeefi and

Scholz, 2017; Pinto et al., 2016; Tappel et al., 2012b, 2012a). These functional PHAs are produced in both native-producing (e.g. Pseudomonas sp.) and recombinant bacteria (e.g. Escherichia coli) from functionalized fatty acid substrates, which enter the β-oxidation pathway and are converted to 3-hydroxyacyl-CoA esters, and subsequently polymerized by a PHA synthase. These functional groups confer significant chemical reactivity to the polymer and are notably amenable to the so- called “Click” reactions, which are desirable due to their selectivity, high yield, stereospecificity, and benign reaction conditions (Kolb et al., 2001). In particular, the strain-promoted azide-alkyne cycloaddition (SPAAC) reaction has seen significant utilization as a conjugation method for

170 nanomedicine, owing to its simple reaction conditions (room temperature, aqueous solvents) and the omission of toxic copper-catalysts, summarized in Scheme 5.1 (Dommerholt et al., 2016).

Scheme 5.1: General strain-promoted alkyne-azide cycloaddition (SPAAC) reaction. For our studies, R1 denotes the alkyl side-chain of azido-PHA monomers and R2 denotes a di-benzyl substitution on the cyclooctyne ring (dibenzylcyclooctyne, DBCO).

The goal of this study was to utilize newly characterized copolymers poly(3- hydroxybutyrate-co-3-hydroxy-10-azidodecanoate) (PHBDN3) and poly(3-hydroxyoctanoate-co-

3-hydroxy-10-azidodecanoate) (PHODN3) to develop a targeted nanoparticle drug delivery system based on folate receptor-mediated endocytosis and the model chemotherapeutic paclitaxel (Fig.

5.1) (Pinto et al., 2016).

Figure 5.1: Proposed synthesis of targeted nanoparticles for drug delivery. Azide-containing PHA copolymers (Azido PHA) will be used to encapsulate the chemotherapeutic agent paclitaxel (PTX, yellow dots). Folic acid analogs will be conjugated to surface-exposed azides (red dots) via strain- promoted azide-alkyne cycloaddition (SPAAC).

171

5.3 Methods

5.3.1 Bulk polymer biosynthesis

Azido-PHAs were biosynthesized as previously described (Pinto et al., 2016). In brief, E. coli RSC02 was transformed with pBBRC1J4SII by heat shock following standard procedures

(Sambrook and Russell, 2001). A 50-mL seed culture (LB-Lennox, composition per liter: 10 g tryptone, 5 g yeast extract, 5 g sodium chloride) supplemented with 50 μg mL-1 kanamycin sulfate was inoculated with LSBJ-pBBRC1J4SII and grown for 16 h at 37°C and 200 rpm rotary shaking.

Growth media (6 × 400 mL LB in 2-L flasks) was supplemented with synthetic 10-azidodecanoic acid (2 mM) and either sodium octanoate or sodium butyrate (8 mM), along with kanamycin sulfate

(50 μg mL-1). The surfactant Brij-35 (4.0 g L-1, Sigma-Aldrich) was added to ensure dissolution of the fatty acids. Growth flasks were inoculated with 0.5% (v/v) seed culture and grown for 48 h at

30°C and 250 rpm. Bacteria were harvested by centrifugation at 3,716 × g for 15 min in 800-mL buckets, washed once with 200-mL 40% ethanol and once with 200-mL Nanopure filtered water.

Cell pellets were transferred to 50-mL centrifuge tubes and resuspended 5-15 mL Nanopure water, frozen, and dried via lyophilization.

Polymer was isolated from dried cells by Soxhlet extraction with 250 mL chloroform

(HPLC Grade, Fisher Scientific) for 6 h, with the extract then concentrated by rotary evaporation.

Polymer was purified via non-solvent precipitation by drop-wise addition to cold methanol (4°C) with moderate stirring, at a 1:10 ratio of chloroform:methanol. Precipitated polymer was collected by centrifugation at 8000 × g and 4°C for 20 min. Polymers were subjected to three rounds of methanol purification, then dried in vacuo and stored at 4°C until needed. Polymer was produced using these procedures as necessary to generate more material.

172

5.3.2 Endotoxin removal and analysis

Growth media for azido PHA endotoxin analysis experiments was composed of; LB, Brij-

35 (4.0 gL-1), kanamycin (50 mg L-1), and the fatty acids 10-azidodecanoic acid (2 mM) and either sodium octanoate or sodium butyrate (8 mM). E. coli RSC02 was transformed with the pBBR-

C1J4SII plasmid by heat shock, and grown on an LB-agar plate at 37°C for 16 hours with kanamycin selection. Individual colonies were picked to inoculate 5 mL LB seed cultures, in triplicate, and grown for 16 h at 37°C and 200 rpm rotary shaking. Seed cultures were used to inoculate triplicate groups of 6 baffled shake flasks each containing 100 mL of growth media at a final seed concentration of 0.5%, for a total of 18 flasks per copolymer. Flasks were cultured for

48 h at 30°C and 250 rpm.

Cells were harvested by centrifugation at 3,716 × g, washed once with 40 mL of 45% ethanol and then once with 40 mL of Nanopure water. Each 600 mL triplicate group was divided in half to have a 3 × 100 mL control group and a 3 × 100 mL sodium hydroxide (NaOH) treatment group. The control groups were resuspended in 5 mL of endotoxin-free water (Fisher Scientific) after harvest, frozen, and dried via lyophilization. The NaOH groups were resuspended in 35 mL of NaOH (0.2 M) in endotoxin-free water and incubated for 5 h at 37°C with 80 rpm reciprocal shaking, in a protocol adapted from Lee et al. (1999). The cell suspension was neutralized with endotoxin-free hydrochloric acid (HCl, 2 M), harvested by centrifugation at 6000 × g for 20 min, and washed once with endotoxin-free water before being frozen and dried via lyophilization.

From this point forward, all glassware was heat-treated at 250° for 1 h to destroy endotoxin, prior to coming in contact with polymer. Polymer from both the control and NaOH treated cells was isolated by Soxhlet extraction (150 mL chloroform, 6 h reflux), then transferred to tared 20- mL scintillation vials. A sample of this crude polymer was set aside (~20 mg) using flame-treated

173 steel implements, with the remainder being purified by non-solvent precipitation in cold methanol

(4°C) with moderate stirring, at a 1:10 ratio of chloroform:methanol. The precipitated polymer was collected by centrifugation at 8000 × g and 4°C for 20 min, re-dissolved in chloroform, and transferred to tared scintillation vials. Each polymer underwent 3 cycles of methanol purification, with 20 mg samples removed each time for endotoxin testing. All samples were dried in vacuo to remove residual solvents before weighing and recorded masses were used for calculations.

Endotoxin testing was conducted using the Pierce™ LAL Chromogenic Endotoxin

Quantitation Kit following the manufacturers recommended procedure. Endotoxin-free water was added to the dried polymer samples at a concentration of 40 μL mg-1 of polymer and incubated at

37°C with 200 rpm rotary shaking for 24 h, adapted from previously described methods (Furrer et al., 2007; Lee et al., 1999). This water was used for the endotoxin assay (50 μL per sample). The absorbance was measured at 410 nm using a Synergy™ H4 microplate reader (BioTek), and endotoxin was quantified as endotoxin units (EU) using a calibration curve of known endotoxin concentrations.

5.3.3 Molecular weight determination

Both weight average (Mw) and number average (Mn) molecular weights were determined using gel permeation chromatography as described previously (Scheel et al., 2016). Polymer samples were dissolved in chloroform to an approximate concentration of 1.0 mg mL-1 and passed through a 0.45 μm PTFE syringe filter. Samples were injected at a volume of 50 μL into a

Shimadzu LC-20AD liquid chromatograph equipped with a Shimadzu SIL-20A autosampler, a

Shimadzu CTO-20A column oven, and a Shimadzu RID-10A refractive index detector. Samples were passed through an 8 × 50 mm styrenedivinylbenzene (S-DVB) guard column (5 μm particles;

Polymer Standards Service) and an 8 × 300 mm S-DVB analytical column (5 μm particles; mixed

174 bed porosity; max molecular weight 1 × 106 Da; Polymer Standards Service product sda083005lim). The column oven was maintained at 40 °C, and chloroform was used as the mobile phase at 1 mL min-1. Polystyrene molecular weight standards with a narrow polydispersity index were employed for calibration. Data analysis was performed using Shimadzu LCSolution software.

5.3.4 Production of PHA nanoparticles

Numerous nanoparticle production strategies were tested using a variety of conditions, including two techniques (single emulsion-solvent evaporation and nanoprecipitation), two different surfactants, seven different organic solvent mixtures, and two different aqueous:organic ratios, which is summarized in Appendix B. The method that produced the smallest average diameter and the most stable nanoparticles (least aggregation) was used for subsequent paclitaxel encapsulation and conjugation experiments; this method is described in detail below.

Polymer solutions of PHODN3 or 50:50 poly(DL-lactide-co-glycolide) (PLGA, purchased from DURECT) were prepared via nanoprecipitation by dissolution in 85:15 acetone:ethanol to a

2.8 mg mL-1 final concentration. The drug paclitaxel (PTX, LC Laboratories) was also dissolved in 85:15 acetone:ethanol to final concentrations of 0.028, 0.14, and 0.28 mg mL-1, corresponding to 1%, 5%, and 10% w/w of the polymer employed. In quadruplicate, the PTX and polymer solutions were combined in equal volumes to produce 20 mL solutions, followed by slow drop- wise addition (~1 mL min-1 by addition funnel) to a beaker containing vigorously stirred deionized water (20 mL; prefiltered through a 0.20 μm PES syringe filter). Paclitaxel-free nanoparticles were produced by combining the polymer solution with pure 85:15 acetone:ethanol and following the same procedure. The resulting solution was left to stir uncovered at room temperature for 24 h.

The resulting aqueous heterogeneous suspension was purified, first through a plug of cotton, and

175 then by gel filtration chromatography (GFC). Aqueous nanoparticle suspensions were loaded (2.5 mL aliquots) onto PD-10 desalting columns containing 8.3 mL of Sephadex G-25 medium (GE

Healthcare Life Sciences) and eluted by gravity with 3.5 mL Nanopure filtered water following manufacturer specifications.

Nanoparticle sizes were determined by dynamic light scattering (DLS) with a Zetasizer

Nano ZS (Malvern) equipped with a 4.0 mW He-Ne laser (λ = 633 nm), operated at room temperature, and at a scattering angle of 173°. Polystyrene cells with 10 mm optical paths were used for DLS experiments, with ~1 mL sample volume and a concentration of polymer nanoparticle suspensions of 0.50–1.0 mg mL-1. Each sample was analyzed in triplicate. Following

DLS analysis, samples were either frozen and lyophilized for paclitaxel quantification or used in subsequent conjugation experiments.

5.3.5 Quantification of paclitaxel by HPLC

High performance liquid chromatography (HPLC) was performed with a Shimadzu

LC2010A HT HPLC system equipped with a reverse phase Zorbax Eclipse XDB-C18 (4.6 x 15 mm, particle size 5 μm, Agilent) analytical column and a UV detector. Two mobile phases were employed, mobile phase A (0.1% formic acid in water) and mobile phase B (0.1% formic acid in acetonitrile). Lyophilized nanoparticle samples were dissolved in 1 mL of acetonitrile, injected (5

μL), and analyzed by isocratic elution consisting of 55% B over 10 min with a flow rate of 1 mL min-1 while monitoring at 227 nm. Column oven was maintained at 40°C. Paclitaxel was quantified using Shimadzu’s LabSolutions LC software based on calibration curves generated with pure paclitaxel. Encapsulation efficiency (EE) was calculated as shown here:

푀푎푠푠 푑푟푢푔 푙표푎푑푒푑 (푚푔) 퐸푛푐푎푝푠푢푙푎푡𝑖표푛 푒푓푓𝑖푐𝑖푒푛푐푦 % (퐸퐸) = ( ) × 100 푀푎푠푠 표푓 푑푟푢푔 푎푑푑푒푑 푡표 푠표푙푢푡𝑖표푛 (푚푔)

176

5.3.6 Conjugation of PEG to PHODN3 and PLGA nanoparticles

Dibenzocyclooctyne-poly(ethylene glycol) (DBCO-PEG, 5kDa, purchased from Nanocs) was conjugated to 5% of the azido moieties on PHODN3 via the SPAAC reaction. The mol% of

1 PHODN3 was calculated by H-NMR as described in Chapter 4. PHA was dissolved in 1.0 mL chloroform at a concentration of 100 mg mL-1 and combined with 2.0 mL of DBCO-PEG solution

(27 mg mL-1). The reaction was carried out at room temperature for 1 h, and successful conjugation was determined by 1H-NMR by the appearance of a multiplet peak at δ 4.5-4.0 corresponding to the two protons alpha to the triazole (Fig. A11, Appendix A). The resulting pegylated-polymer

(PHODN3-PEG) was dried in vacuo and used for nanoparticle production as described above.

PLGA was conjugated to PEG (5 kDa, Sigma Aldrich) as described by El-Gogary et al., with minor adaptations (2014). In brief, PLGA (500 mg, 0.12 mmol), N-hydroxysuccinimide (4 eq, Acros Organics), and N,N`-dicyclohexylcarbodiimide (4 eq, Acros Organics) were dissolved in 1.2 mL dry dichloromethane at room temperature. This solution was filtered (0.45 μm PTFE,

Restek) and added to 1.05 eq PEG and stirred overnight. N,N-diisopropylethylamine (10 eq, Acros

Organics) was added, and the solution was stirred for 48 h. The product was precipitated with cold diethyl ether and collected by centrifugation for 30 min at 3500 rpm. PLGA-PEG was verified by

1H-NMR based on characteristic peaks at δ 5.3, 4.9, and 1.46 (PLGA) and δ 3.3 (PEG), as well as its high solubility in dimethyl sulfoxide (DMSO).

5.4 Results

5.4.1 Endotoxin removal and molecular weight analysis

The endotoxin content of PHBDN3 and PHODN3 was measured using a Limulus amoebocyte lysate chromogenic assay (LAL). Treatment with 0.2M NaOH was effective at removing nearly all endotoxin from both PHBDN3 and PHODN3, reducing the measured

177 endotoxin units (EU) per mg by 96% and 98%, respectively (Fig. 5.2A). Polymer molecular weight

(Mn) remained relatively unchanged between the NaOH and control groups (Fig. 5.2B). Both polymers responded similarly to the treatment; endotoxin was drastically reduced in treated groups and molecular weights were unaffected.

-1 Figure 5.2: (A) Endotoxin present (EU mg ) in crude PHBDN3 (dark gray) and PHODN3 (striped) as determined by the LAL chromogenic endpoint assay. The sodium hydroxide treatment (0.2 M NaOH) removed 0.22 EU mg-1 (~96%) of detectable endotoxin compared to the untreated -1 (Control) PHBDN3, and 0.19 EU mg (~98%) was removed from PHODN3. (B) Average molecular weight (Mn) of crude polymer samples as determined by GPC. Values shown in both (A) and (B) are an average of triplicate experiments, plus or minus the standard deviation around those averages.

The two polymers responded differently to successive methanol purifications. The endotoxin measured from the PHBDN3 control group was reduced by 97% after the first methanol purification down to the same level as the NaOH group, while endotoxin measured from the

PHODN3 control remained unchanged until the final round of purification (Fig. 5.3AB). An

178 increase in endotoxin was observed for PHODN3 after the second round of purification, which is believed to be due to external contamination (Fig. 5.3B). A minor decrease in Mn was observed for both polymers in the NaOH treatment group, which was only observed after methanol purification (Fig. 5.3CD). However, molecular weights increased slightly by the third round of purification for the PHODN3 control and PHBDN3 NaOH groups. No data was collected for the thrice purified PHBDN3 control, as the polymer samples became insoluble in chloroform.

-1 Figure 5.3: Change in endotoxin content (EU mg ) and molecular weight (Mn) over successive methanol (MeOH) purifications for PHBDN3 (A, C) and PHODN3 (B, D). An increase in endotoxin for NaOH-treated PHODN3 (light gray diamonds) was observed after the second methanol purification, believed to be due to contamination from an external source. Endotoxin content of control PHODN3 (dark gray squares) remained high for the first 2 methanol purifications before decreasing dramatically. All values are an average of triplicate experiments, plus or minus the standard deviation around those averages. 179

5.4.2 Comparison of nanoparticle production methods

Numerous methodologies were tested to determine a reproducible production process for uniform nanoparticles <100 nm in diameter. More information on the initial methods used to generate nanoparticles can be found in Appendix B, and a summary of the nanoparticle sizes produced is shown in Fig. 5.4. Initial experiments with PHBDN3 formed large aggregates (data not shown), and the remainder of the experiments were performed exclusively with PHODN3.

Particles produced by single emulsion in sodium dodecyl sulfate (SDS, 4.4 mg mL-1) were much smaller than those produced in poly(vinyl alcohol) (PVA, 4.4 mg mL-1) at 160.7 nm (Fig. 5.4A), and SDS was used in subsequent experiments. Varying the duration of ultrasonication had little effect on particle size (Fig. 5.4B) and reducing the polymer concentration to 15 or 7.6 mg mL-1 only resulted in minor decreases (Fig. 5.4C). Particles produced by the nanoprecipitation method varied widely depending on the organic solvent system used, and most organic solvents resulted in particles of a large range (Fig. 5.4D). Particles were centrifuged and washed to remove residual

SDS, which resulted in the precipitation of larger particles but not smaller ones, depending on the organic solvent used. The acetone/ethanol and dimethylformamide (DMF) solvent systems were chosen for further analysis (Fig. 5.5). Longer centrifugation times enabled the precipitation of larger particles produced using DMF, however the smaller particles remained in suspension, and the acetone/ethanol mixture produced smaller particles overall. Replacing the SDS solution with pure water and using acetone/ethanol as the organic solvent generated the smallest particles yet, and with the narrowest polydispersity. No particles could be precipitated by centrifugation, however the removal of surfactant eliminates the need for a wash step.

180

Figure 5.4: Survey of PHODN3 nanoparticle production methods by size. Particles produced by single emulsion via ultrasonication (A-C) were selected for smallest size using an iterative approach. (A) The surfactants poly(vinyl alcohol) (PVA) and sodium dodecyl sulfate (SDS) were compared first, and SDS selected for subsequent experiments. (B) Effect of varying ultrasonication durations, 50 sec duration used in prior and subsequent experiments. (C) Effect of PHODN3 concentration on size. Particles produced by nanoprecipitation (D) were compared according to organic solvent system and by the ratio of organic solvent to aqueous SDS (4.4 mg mL-1) solution (1:1 or 1:10). Particles in panel D were measured first by DLS after 24 h (SDS, striped) and after pelleting by centrifugation and resuspending in water (Pellet, light gray). Particles suspended in the supernatant after centrifugation were also analyzed (Supernatant, dark gray). Abbreviations: tetrahydrofuran (THF), acetone (ace), ethanol (EtOH), dimethyl sulfoxide (DMSO), dimethylformamide (DMF). Both Ace/DMSO and Ace/EtOH contained 85% acetone (v/v).

181

Figure 5.5: Size comparison of PHODN3 nanoparticles produced by nanoprecipitation. Polymers dissolved in DMF and acetone/ethanol (Ace/EtOH, 85% acetone v/v) were added dropwise to aqueous SDS solutions (4.4 mg mL-1) and centrifuged after 24 h. Particle size was measured by DLS for both the supernatant and the resuspended pellet. Polymers dissolved in acetone/ethanol (Ace/EtOH, 85% acetone v/v) were also added dropwise to pure water, which produced uniform particles measuring 70.6 nm and no pelleted particles. Data shown are the average of triplicate experiments, plus or minus the standard deviation around those averages.

5.4.3 Paclitaxel encapsulation efficiency

PHODN3 and PLGA nanoparticles, along with their pegylated counterparts, were compared with regard to their ability to encapsulate PTX at three loading concentrations (Fig. 5.6).

PHODN3 was far more effective at encapsulating PTX at all three concentrations, and un- pegylated PHODN3 encapsulated approximately 80% of PTX when it was combined at 1% (w/w).

All polymers tested were less effective at encapsulating PTX as the loading concentration increased.

182

Figure 5.6: Encapsulation efficiency of polymeric nanoparticles. Particles were prepared using the nanoprecipitation method and varying concentrations of paclitaxel (PTX). Data shown are the average of quadruplicate experiments, plus or minus the standard deviation around those averages.

5.5 Discussion

The NaOH treatment was very effective at reducing endotoxin toxin levels, bringing these values to below 0.01 EU mg-1, a decrease of approximately 97% compared to the control (Fig.

5.2A). At this concentration of endotoxin, up to 71.4 mg of PHBDN3 or 125 mg of PHODN3 per mL could be administered intravenously under the FDA guidelines of 0.5 EU mL-1, with a maximum of 2.8 g or 5.0 g per dose, respectively (20 EU/device limit) (U.S. Food and Drug

Administration, 2016). However, additional purification is necessary to remove other hydrophobic contaminants such as lipids and proteins, which can be achieved by precipitation in a non-solvent such as methanol (Sevastianov et al., 2003). Precipitation in methanol appeared to remove

- endotoxin quite readily from the PHBDN3 control group, which was reduced to the same EU mg

1 as the NaOH group with a single purification step (Fig. 5.3A). This phenomenon was not observed with PHODN3; the endotoxin was only removed from the control after the final purification (Fig. 5.3B). It is unclear why endotoxin remained associated with PHODN3 through

183 several rounds of purification; it is possible that the lipid A component of LPS formed stronger hydrophobic interactions with the longer side chains of 3-hydoxyoctanoate monomers in comparison with the short side chains of 3-hydroxybutyrate, although this speculation is based solely on the structural characteristics of LPS (Rietschel et al., 1994).

As a polyester, PHAs can be susceptible to backbone cleavage under basic conditions. To determine the effects of our 0.2 M NaOH treatment, we used GPC to analyze the polymer molecular weights with and without treatment (Fig. 5.2B). We found that there was no change in molecular weight observed for the crude polymer after treatment. As the polymers were successively purified by methanol precipitation, the molecular weights of the PHODN3 control group increased incrementally, likely as a result of the loss of lower molecular weight species (Fig.

5.3A). A similar trend was observed for the PHBDN3 NaOH group (Fig. 5.3C). Based on the inconsequential changes observed in the molecular weight post-NaOH treatment, and the efficacy of both methods for removing endotoxin, both purification methods were deemed viable options for low-endotoxin nanoparticle production.

One of the key characteristics of nanoparticle drug-delivery platforms is their size. The circulation time (Ahmed et al., 2005; Hoshyar et al., 2016), rate and mechanism (diffusion, endocytosis, or phagocytosis) of cellular uptake and/or expulsion (Lu et al., 2009; Peruzynska et al., 2015), and intracellular localization (Wu et al., 2019) of nanoparticles have all been shown to be highly dependent on size, with differences observed for sizes in a range as small as 20 nm.

Optimal size depends on the target cell type and the desired route of internalization, however most of these studies concluded that nanoparticles of approximately 90-100 nm show desirable internalization efficiency. Over numerous iterative experiments comparing nanoparticle production methods, we were able to determine the ideal conditions for producing PHODN3

184 nanoparticles of a consistent, uniform size (Fig. 5.4 & 5.5). Nanoprecipitation of our polymer into pure water not only resulted in a desirable particle size, it also eliminated time-consuming surfactant removal steps and the use of poorly scalable techniques such as ultrasonication and centrifugation.

The PHODN3 far outperformed the PLGA, regarded as the gold standard for polymeric nanoparticles, in its ability to encapsulate paclitaxel in this study (Fig. 5.6). PLGA has been shown to be capable of higher efficiency with respect to PTX encapsulation, with efficiencies of approximately 70-75% reported (Betancourt et al., 2009; Danhier et al., 2009; Wang et al., 2011).

These studies differed in the techniques used to prepare nanoparticles (single emulsion, double emulsion, and nanoprecipitation), the organic and aqueous solutions employed, and the use of polymer blends, demonstrating that drug encapsulation is dependent on multiple factors. These studies also produced nanoparticles larger than the sizes achieved here with PHODN3, ranging from 112-240 nm, which as described above is an important factor in cell uptake. Further characterization of our nanoparticles with regard to cellular uptake and drug release is needed to determine their efficacy.

The data presented in this chapter is largely preliminary, setting the foundation for a new polymeric drug delivery system. The next steps needed are to evaluate the shelf-life stability of

PTX-loaded PHODN3 nanoparticles and the release of PTX over time, the in vitro cellular uptake kinetics of folate-conjugated nanoparticles by human cancer cells, and the cytotoxicity of these nanoparticles. Currently, work is underway assessing cellular uptake of folate-labelled nanoparticles by human breast cancer cells, which will be analyzed by live-culture fluorescence microscopy (IncuCyte®, Sartorius) with fluorophore-conjugated nanoparticles. Once cell uptake kinetics have been established, cell viability after exposure to paclitaxel-loaded nanoparticles will

185 be assessed to determine the cytotoxicity, and these experiments are in the preliminary stages of study design.

5.6 References

Ahmed, M., Lukyanov, A.N., Torchilin, V., Tournier, H., Schneider, A.N., Goldberg, S.N., 2005.

Combined Radiofrequency Ablation and Adjuvant Liposomal Chemotherapy: Effect of

Chemotherapeutic Agent, Nanoparticle Size, and Circulation Time. Journal of Vascular

and Interventional Radiology 16, 1365–1371.

https://doi.org/10.1097/01.RVI.0000175324.63304.25

Bahrami, B., Mohammadnia-Afrouzi, M., Bakhshaei, P., Yazdani, Y., Ghalamfarsa, G., Yousefi,

M., Sadreddini, S., Jadidi-Niaragh, F., Hojjat-Farsangi, M., 2015. Folate-conjugated

nanoparticles as a potent therapeutic approach in targeted cancer therapy. Tumour Biol. 36,

5727–5742. https://doi.org/10.1007/s13277-015-3706-6

Betancourt, T., Byrne, J.D., Sunaryo, N., Crowder, S.W., Kadapakkam, M., Patel, S., Casciato, S.,

Brannon-Peppas, L., 2009. PEGylation strategies for active targeting of PLA/PLGA

nanoparticles. J Biomed Mater Res A 91, 263–276. https://doi.org/10.1002/jbm.a.32247

Calvano, S.E., Coyle, S.M., 2012. Experimental human endotoxemia: a model of the systemic

inflammatory response syndrome? Surg Infect (Larchmt) 13, 293–299.

https://doi.org/10.1089/sur.2012.155

Danhier, F., Ansorena, E., Silva, J.M., Coco, R., Le Breton, A., Préat, V., 2012. PLGA-based

nanoparticles: an overview of biomedical applications. J Control Release 161, 505–522.

https://doi.org/10.1016/j.jconrel.2012.01.043

Danhier, F., Lecouturier, N., Vroman, B., Jérôme, C., Marchand-Brynaert, J., Feron, O., Préat, V.,

2009. Paclitaxel-loaded PEGylated PLGA-based nanoparticles: In vitro and in vivo

186

evaluation. Journal of Controlled Release 133, 11–17.

https://doi.org/10.1016/j.jconrel.2008.09.086

Dommerholt, J., Rutjes, F.P.J.T., van Delft, F.L., 2016. Strain-Promoted 1,3-Dipolar

Cycloaddition of Cycloalkynes and Organic Azides. Top Curr Chem (Z) 374, 16.

https://doi.org/10.1007/s41061-016-0016-4

El-Gogary, R.I., Rubio, N., Wang, J.T.-W., Al-Jamal, W.T., Bourgognon, M., Kafa, H., Naeem,

M., Klippstein, R., Abbate, V., Leroux, F., Bals, S., Van Tendeloo, G., Kamel, A.O., Awad,

G.A.S., Mortada, N.D., Al-Jamal, K.T., 2014. Polyethylene Glycol Conjugated Polymeric

Nanocapsules for Targeted Delivery of Quercetin to Folate-Expressing Cancer Cells in

Vitro and in Vivo. ACS Nano 8, 1384–1401. https://doi.org/10.1021/nn405155b

Foraker, A.B., Khantwal, C.M., Swaan, P.W., 2003. Current perspectives on the cellular uptake

and trafficking of riboflavin. Adv. Drug Deliv. Rev. 55, 1467–1483.

https://doi.org/10.1016/j.addr.2003.07.005

Furrer, P., Panke, S., Zinn, M., 2007. Efficient recovery of low endotoxin medium-chain-length

poly([R]-3-hydroxyalkanoate) from bacterial biomass. Journal of Microbiological Methods

69, 206–213. https://doi.org/10.1016/j.mimet.2007.01.002

Hadjipanayis, C.G., Machaidze, R., Kaluzova, M., Wang, L., Schuette, A.J., Chen, H., Wu, X.,

Mao, H., 2010. EGFRvIII antibody-conjugated iron oxide nanoparticles for magnetic

resonance imaging-guided convection-enhanced delivery and targeted therapy of

glioblastoma. Cancer Res. 70, 6303–6312. https://doi.org/10.1158/0008-5472.CAN-10-

1022

187

Hans, M.L., Lowman, A.M., 2002. Biodegradable nanoparticles for drug delivery and targeting.

Current Opinion in Solid State and Materials Science 6, 319–327.

https://doi.org/10.1016/S1359-0286(02)00117-1

Hillaireau, H., Couvreur, P., 2009. Nanocarriers’ entry into the cell: relevance to drug delivery.

Cell. Mol. Life Sci. 66, 2873–2896. https://doi.org/10.1007/s00018-009-0053-z

Hoshyar, N., Gray, S., Han, H., Bao, G., 2016. The effect of nanoparticle size on in vivo

pharmacokinetics and cellular interaction. Nanomedicine 11, 673–692.

https://doi.org/10.2217/nnm.16.5

Kennedy, M.D., Jallad, K.N., Lu, J., Low, P.S., Ben-Amotz, D., 2003. Evaluation of folate

conjugate uptake and transport by the choroid plexus of mice. Pharm. Res. 20, 714–719.

Kolb, H.C., Finn, M.G., Sharpless, K.B., 2001. Click Chemistry: Diverse Chemical Function from

a Few Good Reactions. Angewandte Chemie International Edition 40, 2004–2021.

https://doi.org/10.1002/1521-3773(20010601)40:11<2004::AID-ANIE2004>3.0.CO;2-5

Lee, M.S., Dees, E.C., Wang, A.Z., 2017. Nanoparticle-Delivered Chemotherapy: Old Drugs in

New Packages. Oncology (Williston Park, N.Y.) 31, 198–208.

Lee, S.Y., Choi, J., Han, K., Song, J.Y., 1999. Removal of endotoxin during purification of poly

(3-hydroxybutyrate) from gram-negative bacteria. Applied and environmental

microbiology 65, 2762–2764.

Ling, Y., Wei, K., Luo, Y., Gao, X., Zhong, S., 2011. Dual docetaxel/superparamagnetic iron oxide

loaded nanoparticles for both targeting magnetic resonance imaging and cancer therapy.

Biomaterials 32, 7139–7150. https://doi.org/10.1016/j.biomaterials.2011.05.089

188

Lu, F., Wu, S.-H., Hung, Y., Mou, C.-Y., 2009. Size effect on cell uptake in well-suspended,

uniform mesoporous silica nanoparticles. Small 5, 1408–1413.

https://doi.org/10.1002/smll.200900005

Nkrumah-Agyeefi, S., Scholz, C., 2017. Chemical modification of functionalized

polyhydroxyalkanoates via “Click” chemistry: A proof of concept. International journal of

biological macromolecules. https://doi.org/10.1016/j.ijbiomac.2016.11.118

Paulos, C.M., Reddy, J.A., Leamon, C.P., Turk, M.J., Low, P.S., 2004. Ligand Binding and

Kinetics of Folate Receptor Recycling in Vivo: Impact on Receptor-Mediated Drug

Delivery. Mol Pharmacol 66, 1406–1414. https://doi.org/10.1124/mol.104.003723

Peruzynska, M., Cendrowski, K., Barylak, M., Roginska, D., Tarnowski, M., Tkacz, M.,

Kurzawski, M., Machalinski, B., Mijowska, E., Drozdzik, M., 2015. Study on size effect

of the silica nanospheres with solid core and mesoporous shell on cellular uptake. Biomed

Mater 10, 065012. https://doi.org/10.1088/1748-6041/10/6/065012

Petrelli, F., Borgonovo, K., Barni, S., 2010. Targeted delivery for breast cancer therapy: the history

of nanoparticle-albumin-bound paclitaxel. Expert Opin Pharmacother 11, 1413–1432.

https://doi.org/10.1517/14656561003796562

Pinto, A., Ciesla, J.H., Palucci, A., Sutliff, B.P., Nomura, C.T., 2016. Chemically Intractable No

More: In Vivo Incorporation of “Click”-Ready Fatty Acids into Poly-[(R)-3-

hydroxyalkanoates] in Escherichia coli. ACS Macro Letters 5, 215–219.

https://doi.org/10.1021/acsmacrolett.5b00823

Poletto, F.S., Fiel, L.A., Donida, B., Ré, M.I., Guterres, S.S., Pohlmann, A.R., 2008. Controlling

the size of poly(hydroxybutyrate-co-hydroxyvalerate) nanoparticles prepared by

emulsification–diffusion technique using ethanol as surface agent. Colloids and Surfaces

189

A: Physicochemical and Engineering Aspects 324, 105–112.

https://doi.org/10.1016/j.colsurfa.2008.04.003

Pramual, S., Assavanig, A., Bergkvist, M., Batt, C.A., Sunintaboon, P., Lirdprapamongkol, K.,

Svasti, J., Niamsiri, N., 2016. Development and characterization of bio-derived

polyhydroxyalkanoate nanoparticles as a delivery system for hydrophobic photodynamic

therapy agents. J Mater Sci Mater Med 27, 40. https://doi.org/10.1007/s10856-015-5655-4

Rai, R., Keshavarz, T., Roether, J.A., Boccaccini, A.R., Roy, I., 2011. Medium chain length

polyhydroxyalkanoates, promising new biomedical materials for the future. Materials

Science and Engineering: R: Reports 72, 29–47.

https://doi.org/10.1016/j.mser.2010.11.002

Rietschel, E.T., Kirikae, T., Schade, F.U., Mamat, U., Schmidt, G., Loppnow, H., Ulmer, A.J.,

Zähringer, U., Seydel, U., Di Padova, F., 1994. Bacterial endotoxin: molecular

relationships of structure to activity and function. The FASEB Journal 8, 217–225.

https://doi.org/10.1096/fasebj.8.2.8119492

Sambrook, J.J., Russell, D.W., 2001. Molecular cloning: A laboratory manual. Cold Spring Harbor

Laboratory Press, Cold Spring Harbor, NY.

Sarris, A.H., Hagemeister, F., Romaguera, J., Rodriguez, M.A., McLaughlin, P., Tsimberidou,

A.M., Medeiros, L.J., Samuels, B., Pate, O., Oholendt, M., Kantarjian, H., Burge, C.,

Cabanillas, F., 2000. Liposomal vincristine in relapsed non-Hodgkin’s lymphomas: early

results of an ongoing phase II trial. Ann. Oncol. 11, 69–72.

https://doi.org/10.1023/a:1008348010437

190

Scheel, R.A., Ji, L., Lundgren, B.R., Nomura, C.T., 2016. Enhancing poly(3-hydroxyalkanoate)

production in Escherichia coli by the removal of the regulatory gene arcA. AMB Express

6, 120. https://doi.org/10.1186/s13568-016-0291-z

Sevastianov, V.I., Perova, N.V., Shishatskaya, E.I., Kalacheva, G.S., Volova, T.G., 2003.

Production of purified polyhydroxyalkanoates (PHAs) for applications in contact with

blood. Journal of Biomaterials Science, Polymer Edition 14, 1029–1042.

https://doi.org/10.1163/156856203769231547

Tappel, R.C., Kucharski, J.M., Mastroianni, J.M., Stipanovic, A.J., Nomura, C.T., 2012a.

Biosynthesis of Poly[(R)-3-hydroxyalkanoate] Copolymers with Controlled Repeating

Unit Compositions and Physical Properties. Biomacromolecules 13, 2964.

Tappel, R.C., Wang, Q., Nomura, C.T., 2012b. Precise control of repeating unit composition in

biodegradable poly(3-hydroxyalkanoate) polymers synthesized by Escherichia coli.

Journal of Bioscience and Bioengineering 113, 480–486.

https://doi.org/10.1016/j.jbiosc.2011.12.004

U.S. Food and Drug Administration, 2016. Submission and Review of Sterility Information in

Premarket Notification (510(k)) Submissions for Devices Labeled as Sterile - Guidance for

Industry and Food and Drug Administration Staff.

Wang, H., Zhao, Y., Wu, Y., Hu, Y., Nan, K., Nie, G., Chen, H., 2011. Enhanced anti-tumor

efficacy by co-delivery of doxorubicin and paclitaxel with amphiphilic methoxy PEG-

PLGA copolymer nanoparticles. Biomaterials 32, 8281–8290.

https://doi.org/10.1016/j.biomaterials.2011.07.032

191

Wu, M., Guo, H., Liu, L., Liu, Y., Xie, L., 2019. Size-dependent cellular uptake and localization

profiles of silver nanoparticles. Int J Nanomedicine 14, 4247–4259.

https://doi.org/10.2147/IJN.S201107

Xiong, Y.-C., Yao, Y.-C., Zhan, X.-Y., Chen, G.-Q., 2010. Application of polyhydroxyalkanoates

nanoparticles as intracellular sustained drug-release vectors. J Biomater Sci Polym Ed 21,

127–140. https://doi.org/10.1163/156856209X410283

Yu, M.K., Park, J., Jon, S., 2012. Targeting Strategies for Multifunctional Nanoparticles in Cancer

Imaging and Therapy. Theranostics 2, 3–44. https://doi.org/10.7150/thno.3463

Zhao, R., Diop-Bove, N., Goldman, I.D., 2014. Enhanced Receptor-Mediated Endocytosis and

Cytotoxicity of a Folic Acid-Desacetylvinblastine Monohydrazide Conjugate in a

Pemetrexed-Resistant Cell Line Lacking Folate-Specific Facilitative Carriers but with

Increased Folate Receptor Expression. Mol Pharmacol 85, 310–321.

https://doi.org/10.1124/mol.113.089110

192

Chapter 6: Conclusions and future outlook

6.1 Discussion

The most significant finding of this dissertation is that Escherichia coli LSBJ is able to produce homopolymeric MCL PHAs from fatty acid substrate at yields approaching the theoretical maximum using an optimized fed-batch, high-density fermentation process. Although other recent works using engineered Pseudomonas putida KT2440 have achieved similar yields for near- homopolymeric PHD (highest to date; 11.8 g L-1 , productivity of 0.41 g L-1 h-1), none have succeeded in expanding this compositional control to other MCL PHAs (Gao et al., 2018; Oliveira et al., 2020). Furthermore, no other described processes have successfully produced PHAs from functionalized fatty acids (10-undecenoic acid, 10-bromodecanoic, and 10-azidodecanoic) at the yields demonstrated in Chapter 4. This has the considerable potential to expand the production of

PHAs for biomedical materials, which require tightly controlled material properties and can benefit greatly from chemically reactive moieties (Hazer et al., 2012; Kai and Loh, 2014; Rai et al., 2011).

Removal of the arcA transcriptional regulator from E. coli LSBJ was shown to enhance both SCL and MCL PHA biosynthesis from the respective fatty acid substrates (Chapter 2). This is believed to have occurred due to the increased expression of the β-oxidation genes fadD, fadE, and fadL in the mutant strain, which saw relative increases of 2-fold, 31-fold, and 10-fold compared to E. coli LSBJ, respectively (Scheel et al., 2016). Interestingly, the expression of the atoDAEB operon (implicated in SCL fatty acid uptake and activation) was greatly upregulated despite no previously reported interactions between the ArcA and Ato systems (Cho et al., 2006;

Theodorou et al., 2011, 2006). These results suggest that ArcA-mediated transcriptional regulation plays a role in PHA biosynthesis in E. coli, and that crosstalk between ArcA and SCL fatty acid

193 metabolism may be more complex than previously assumed. Removal of the OmpR regulator was also shown to deleteriously affect PHA biosynthesis, likely due to a toxic phenotype.

The use of lignocellulosic waste substrates as bacterial feedstock has experienced significant interest in recent years due to growing concerns about environmental sustainability

(H. Isikgor and Remzi Becer, 2015; Zhang, 2008). One such substrate, enzymatically hydrolyzed paper mill waste fines, was a surprisingly effective feedstock for the biosynthesis of PHB and

PHB-co-MCL PHAs in E. coli LSBJ, discussed in Chapter 3 (Min et al., 2015; Scheel et al., 2019).

Both PHB and PHB-co-MCL PHA yields increased by approximately 100% when pure sugar substrates were replaced with hydrolysate, although the MCL monomer content was reduced.

Progress towards the nanoparticle drug delivery system discussed in Chapter 5 is still underway, with promising preliminary results presented therein. The field of nanomedicine, of which nanoparticle drug delivery is a subset, is rapidly expanding but limited in scope to a narrow selection of materials including PLGA and PEG (Danhier et al., 2012; Lee et al., 2017; Santander-

Ortega et al., 2010). Previous studies investigating PHB and PHBV nanoparticles found them to be inferior materials compared to PLGA and PLGA-based blends, which is why the objective of

Chapter 5 was to produce superior particles using functional MCL PHAs (Pinto et al., 2016; Poletto et al., 2008; Pramual et al., 2016). Future experiments are needed to evaluate the shelf-life stability of PTX-loaded PHODN3 nanoparticles and the release of PTX over time, the in vitro cellular uptake kinetics of folate-conjugated nanoparticles by human cancer cells, and the cytotoxicity of these nanoparticles. The results of these experiments will ultimately determine the efficacy of this novel drug-delivery platform.

Limitations in the production of MCL PHAs are largely due to expensive feed stocks, limited bioprocess productivity, and difficulty in controlling the monomeric composition to

194 achieve desired material properties (Blunt et al., 2019; Hazer et al., 2012; Rai et al., 2011). The work presented here has made some advancements in this field, although the current dependence on expensive fatty acid substrates limits the applicability to higher-performance biomedical materials rather than more common consumer plastics (Kai and Loh, 2014; Sparks and Scholz,

2009). Further modification to the bioprocess described in Chapter 4 to allow the production of

SCL monomers such as 3HB or 4HB from glycolytic metabolites, while simultaneously incorporating desirable MCL monomers from fatty acid substrates, would go a long way towards expanding the availability of these novel materials.

6.2 References

Blunt, W., Dartiailh, C., Sparling, R., Gapes, D.J., Levin, D.B., Cicek, N., 2019. Development of

High Cell Density Cultivation Strategies for Improved Medium Chain Length

Polyhydroxyalkanoate Productivity Using Pseudomonas putida LS46. Bioengineering

(Basel) 6. https://doi.org/10.3390/bioengineering6040089

Cho, B.K., Knight, E.M., Palsson, B.O., 2006. Transcriptional regulation of the fad regulon genes

of Escherichia coli by ArcA. Microbiology 152, 2207–2219. https://doi.org/152/8/2207

[pii]

Danhier, F., Ansorena, E., Silva, J.M., Coco, R., Le Breton, A., Préat, V., 2012. PLGA-based

nanoparticles: an overview of biomedical applications. J Control Release 161, 505–522.

https://doi.org/10.1016/j.jconrel.2012.01.043

Gao, J., Vo, M.T., Ramsay, J.A., Ramsay, B.A., 2018. Overproduction of MCL-PHA with high 3-

hydroxydecanoate Content. Biotechnology and Bioengineering 115, 390–400.

https://doi.org/10.1002/bit.26474

195

Hazer, D.B., Kılıçay, E., Hazer, B., 2012. Poly(3-hydroxyalkanoate)s: Diversification and

biomedical applications. Materials Science and Engineering: C 32, 637–647.

https://doi.org/10.1016/j.msec.2012.01.021

H. Isikgor, F., Remzi Becer, C., 2015. Lignocellulosic biomass: a sustainable platform for the

production of bio-based chemicals and polymers. Polymer Chemistry 6, 4497–4559.

https://doi.org/10.1039/C5PY00263J

Kai, D., Loh, X.J., 2014. Polyhydroxyalkanoates: Chemical Modifications Toward Biomedical

Applications. ACS Sustainable Chem. Eng. 2, 106–119.

https://doi.org/10.1021/sc400340p

Lee, M.S., Dees, E.C., Wang, A.Z., 2017. Nanoparticle-Delivered Chemotherapy: Old Drugs in

New Packages. Oncology (Williston Park, N.Y.) 31, 198–208.

Min, B.C., Bhayani, B.V., Jampana, V.S., Ramarao, B.V., 2015. Enhancement of the enzymatic

hydrolysis of fines from recycled paper mill waste rejects. Bioresources and Bioprocessing

2, 40. https://doi.org/10.1186/s40643-015-0068-2

Oliveira, G.H.D., Zaiat, M., Rodrigues, J.A.D., Ramsay, J.A., Ramsay, B.A., 2020. Towards the

Production of mcl-PHA with Enriched Dominant Monomer Content: Process Development

for the Sugarcane Biorefinery Context. J Polym Environ 28, 844–853.

https://doi.org/10.1007/s10924-019-01637-2

Pinto, A., Ciesla, J.H., Palucci, A., Sutliff, B.P., Nomura, C.T., 2016. Chemically Intractable No

More: In Vivo Incorporation of “Click”-Ready Fatty Acids into Poly-[(R)-3-

hydroxyalkanoates] in Escherichia coli. ACS Macro Letters 5, 215–219.

https://doi.org/10.1021/acsmacrolett.5b00823

196

Poletto, F.S., Fiel, L.A., Donida, B., Ré, M.I., Guterres, S.S., Pohlmann, A.R., 2008. Controlling

the size of poly(hydroxybutyrate-co-hydroxyvalerate) nanoparticles prepared by

emulsification–diffusion technique using ethanol as surface agent. Colloids and Surfaces

A: Physicochemical and Engineering Aspects 324, 105–112.

https://doi.org/10.1016/j.colsurfa.2008.04.003

Pramual, S., Assavanig, A., Bergkvist, M., Batt, C.A., Sunintaboon, P., Lirdprapamongkol, K.,

Svasti, J., Niamsiri, N., 2016. Development and characterization of bio-derived

polyhydroxyalkanoate nanoparticles as a delivery system for hydrophobic photodynamic

therapy agents. J Mater Sci Mater Med 27, 40. https://doi.org/10.1007/s10856-015-5655-4

Rai, R., Keshavarz, T., Roether, J.A., Boccaccini, A.R., Roy, I., 2011. Medium chain length

polyhydroxyalkanoates, promising new biomedical materials for the future. Materials

Science and Engineering: R: Reports 72, 29–47.

https://doi.org/10.1016/j.mser.2010.11.002

Santander-Ortega, M.J., Csaba, N., González, L., Bastos-González, D., Ortega-Vinuesa, J.L.,

Alonso, M.J., 2010. Protein-loaded PLGA–PEO blend nanoparticles: encapsulation,

release and degradation characteristics. Colloid Polym Sci 288, 141–150.

https://doi.org/10.1007/s00396-009-2131-z

Scheel, R.A., Fusi, A.D., Min, B.C., Thomas, C.M., Ramarao, B.V., Nomura, C.T., 2019.

Increased Production of the Value-Added Biopolymers Poly(R-3-Hydroxyalkanoate) and

Poly(γ-Glutamic Acid) From Hydrolyzed Paper Recycling Waste Fines. Front. Bioeng.

Biotechnol. 7. https://doi.org/10.3389/fbioe.2019.00409

197

Scheel, R.A., Ji, L., Lundgren, B.R., Nomura, C.T., 2016. Enhancing poly(3-hydroxyalkanoate)

production in Escherichia coli by the removal of the regulatory gene arcA. AMB Express

6, 120. https://doi.org/10.1186/s13568-016-0291-z

Sparks, J., Scholz, C., 2009. Evaluation of a cationic poly(β-hydroxyalkanoate) as a plasmid DNA

delivery system. Biomacromolecules 10, 1715–1719. https://doi.org/10.1021/bm900372x

Theodorou, E.C., Theodorou, M.C., Samali, M.N., Kyriakidis, D.A., 2011. Activation of the

AtoSC two-component system in the absence of the AtoC N-terminal receiver domain in

E. coli. Amino Acids 40, 421–430. https://doi.org/10.1007/s00726-010-0652-x

Theodorou, M.C., Panagiotidis, C.A., Panagiotidis, C.H., Pantazaki, A.A., Kyriakidis, D.A., 2006.

Involvement of the AtoS-AtoC signal transduction system in poly-(R)-3-hydroxybutyrate

biosynthesis in Escherichia coli. Biochim. Biophys. Acta 1760, 896–906.

https://doi.org/10.1016/j.bbagen.2006.01.020

Zhang, Y.-H.P., 2008. Reviving the carbohydrate economy via multi-product lignocellulose

biorefineries. J Ind Microbiol Biotechnol 35, 367–375. https://doi.org/10.1007/s10295-

007-0293-6

198

Appendix A: Nuclear Magnetic Resonance Spectroscopy Data

The data presented in this appendix are from nuclear magnetic resonance (NMR) experiments performed to characterize PHAs and other small molecules as described in each appropriate chapter. Unless otherwise specified, 1H NMR spectra were recorded on a Bruker AVANCE III

600 MHz instrument, and were calibrated using residual undeuterated solvents as internal reference (chloroform, δ = 7.26 ppm; water, δ = 4.79 ppm). Chemical shifts (δ) are reported in parts per million (ppm); NMR peak multiplicities are denoted by the following abbreviations: s = singlet, d = doublet, t = triplet, q = quartet, p = pentet, sext = sextet, dd = doublet of doublets, dt = doublet of triplets, m = multiplet, br = broad. Spectra were processed with Bruker TopSpin v3.5pI2.

3 2 4

Fig. A1: 1H-NMR (600 MHz, CDCl3) spectrum of poly(R-3-hydroxybutyrate) (PHB) extracted from E. coli RSC02; δ 5.28-5.23 (sext, 1H), 2.62-2.45 (m, 2H), 1.28-1.25 (d, 3H).

199

*

3 2 4 5

Fig. A.2: 1H-NMR (600 MHz, CDCl3) spectrum of poly(R-3-hydroxyvalerate) extracted from E. coli RSC02; δ 5.17-5.13 (p, 1H), 2.59-2.50 (m, 2H), 1.66-1.59 (m, 2H), 0.91-0.88 (t, 3H). The asterisk (*) at δ 1.53 denotes a water impurity (Fulmer et al., 2010).

*

3 2 4 5 6

Fig. A3: 1H-NMR (600 MHz, CDCl3) spectrum of poly(R-3-hydroxyhexanoate) extracted from E. coli RSC02; δ 5.21-5.18 (p, 1H), 2.59-2.48 (m, 2H), 1.60-1.54 (m, 2H), 1.35-1.27 (m, 2H), 0.91 (t, 3H). The asterisk (*) at δ 1.53 denotes a water impurity (Fulmer et al., 2010).

200

Fig. A4: 1H-NMR (600 MHz, 50% D2O) of crude hydrolysate in a 1:10 dilution with glucosamine as an internal standard; δ 5.45 (d, glucosamine α-anomeric H), δ 5.24-5.23 (d, glucose α-anomeric H), δ 5.20-5.19 (d, xylose α-anomeric proton), δ 4.95-4.94 (d, glucosamine β-anomeric H), δ 4.65- 4.64 (d, glucose β-anomeric H), δ 4.57 (d, xylose β-anomeric H), δ 1.97 (s, 3H from acetate).

4 PGA

2 3

Fig. A5: 1H-NMR (600 MHz, D2O) of poly(γ-glutamic acid) (PGA); δ 4.41 (m, 1H), δ 3.08-2.88 (dd, unknown H), δ 2.46 (m, 2H), δ 2.11-1.99 (doublet of broad multiplets, 2H).

201

Current Data Parameters NAME RAS20170123 ppm EXPNO 4 PROCNO 1

F2 - Acquisition Parameters Date_ 20170123 Time 15.56 INSTRUM spect PROBHD 5 mm CPPBBO BB 2.0 PULPROG cosygpppqf TD 2048 SOLVENT D2O NS 4 DS 8 SWH 3184.713 Hz 2.5 FIDRES 1.555036 Hz AQ 0.3215360 sec RG 29.09 DW 157.000 usec DE 10.00 usec TE 303.2 K 3.0 D0 0.00000300 sec D1 1.80625904 sec D11 0.03000000 sec D12 0.00002000 sec D13 0.00000400 sec D16 0.00020000 sec 3.5 IN0 0.00031400 sec

======CHANNEL f1 ======SFO1 600.1024436 MHz NUC1 1H 4.0 P0 10.60 usec P1 10.60 usec P17 2500.00 usec PLW1 28.00000000 W PLW10 4.65399981 W 4.5 ======GRADIENT CHANNEL ===== GPNAM[1] SMSQ10.100 GPZ1 10.00 % P16 1000.00 usec

F1 - Acquisition parameters 5.0 TD 128 SFO1 600.1024 MHz FIDRES 49.761147 Hz SW 5.307 ppm FnMODE QF

5.5 F2 - Processing parameters SI 1024 SF 600.1000000 MHz WDW QSINE SSB 0 6.0 LB 0 Hz GB 0 PC 1.40

F1 - Processing parameters SI 1024 6.5 MC2 QF SF 600.1000000 MHz WDW QSINE SSB 0 LB 0 Hz 6.5 6.0 5.5 5.0 4.5 4.0 3.5 3.0 2.5 2.0 ppm GB 0 Fig. A6: COSY spectra (600 MHz, D2O) of poly(γ-glutamic acid). Determined that the unknown shift at δ 3.08-2.88 (dd, Fig. A.5) was not connected to the polymer.

202

Fig. A7: 1H-NMR (600 MHz, CDCl3) spectrum of poly(R-3-hydroxy-10-undecenoate) extracted from E. coli LSBJ after bioreactor fermentation; δ 5.83-5.76 (m, 1H), 5.21-5.18 (p, 1H), 5.00-4.92 (dd, 2H), 2.59-2.48 (m, 2H), 2.05-2.01 (q, 2H), 1.60-1.54 (m, 2H), 1.38-1.28 (m, 8H).

Fig. A8: 1H-NMR (600 MHz, CDCl3) spectrum of poly(R-3-hydroxy-10-bromodecanoate) extracted from E. coli LSBJ after bioreactor fermentation; δ 5.21-5.18 (p, 1H), 3.37 (t, 2H), 2.59- 2.48 (m, 2H), 1.84-1.82 (p, 2H), 1.60-1.54 (m, 2H), 1.35-1.27 (m, 2H), 1.28 (m, 6H). The asterisk (*) at δ 3.43 denotes a methanol impurity (Fulmer et al., 2010).

203

Fig. A9: 1H-NMR (800 MHz, CDCl3) spectrum of poly(R-3-hydroxyoctanoate-co-3-hydroxy-10- azidodecanoate) extracted from E. coli LSBJ after 24 h bioreactor fermentation; δ 5.19-5.17 (p, 1H), 3.26-3.24 (t, 1H), 2.58-2.48 (m, 2H), 1.59-1.58 (m, 2H), 1.36-1.25 (m, 6H/8H), 0.89-0.87 (t, 3H).

Fig. A10: 1H-NMR (800 MHz, CDCl3) spectrum of poly(R-3-hydroxyoctanoate-co-3-hydroxy- 10-azidodecanoate) extracted from E. coli LSBJ after 48 h bioreactor fermentation; δ 5.19-5.17 (p, 1H), 3.26-3.24 (t, 1H), 2.58-2.48 (m, 2H), 1.59-1.58 (m, 2H), 1.36-1.25 (m, 6H/8H), 0.89-0.87 (t, 3H).

204

7.43

7.28

7.26

7.08

5.24

5.23

5.22

5.19

5.18

5.17

5.16

5.15

3.91

3.76

3.75

3.74

3.68

3.65

3.63

3.59

3.55

3.54

3.54

3.53

3.52

3.51

3.51

3.37

3.26

3.24

3.23

2.63

2.62

2.61

2.59

2.58

2.56

2.55

2.54

2.52

2.51

2.50

2.48

2.47

2.45

2.44

2.21

2.00

1.99

1.78

1.67

1.58

1.57

1.56

1.49

1.47

1.46

1.42

1.39

1.35

1.34 1.29

4.50 4.45 4.40 4.35 4.30 4.25 ppm

7.5 7.0 6.5 6.0 5.5 5.0 4.5 4.0 3.5 3.0 2.5 2.0 1.5 1.0 ppm

0.04

0.04

1.30

0.02

5.62

0.49

2.68

3.04 9.13 3.00 Fig. A11: 1H-NMR (600 MHz, DMSO-d6) spectrum of poly(R-3-hydroxyoctanoate-co-3- hydroxy-10-azidodecanoate)-graft-dibenzocyclooctyne-poly(ethylene glycol methyl ether), prepared by strain-promoted azide-alkyne cycloaddition as described in Chapter 5.

References Fulmer, G.R., Miller, A.J.M., Sherden, N.H., Gottlieb, H.E., Nudelman, A., Stoltz, B.M., Bercaw,

J.E., Goldberg, K.I., 2010. NMR Chemical Shifts of Trace Impurities: Common

Laboratory Solvents, Organics, and Gases in Deuterated Solvents Relevant to the

Organometallic Chemist. Organometallics 29, 2176–2179.

https://doi.org/10.1021/om100106e

205

Appendix B: Nanoparticle Methods Development

This appendix contains supplemental information on the methodologies used to produce nanoparticles in Figure 5.4 and Figure 5.5.

Nanoparticle production by single emulsion

Nanoparticles produced for Figure 5.4ABC were produced using a single emulsion technique adapted from (Budhian et al., 2007; Mainardes and Evangelista, 2005). Poly[(R)-3- hydroxyoctanoate-co-(R)-3-hydroxy-10-azidodecanoate] (PHODN3) was dissolved in 0.5 mL chloroform, then added to 1 mL of surfactant solution (4.44 mg mL-1, sodium dodecylsulfate or poly[vinyl alcohol]) while continuously mixing vigorously by vortex. This solution was then sonicated using a Branson Analog Sonifier 250A fitted with a 1/8” microtip at an output of 4 for

50 seconds (unless otherwise noted), with 10 second on/off pulses. The test tube containing the emulsion was kept in an ice bath during the off cycles.

The sonicated emulsion solution was transferred to 45 mL of surfactant solution (SDS or

PVA, the same as the initial solution) and stirred for 3 hours to evaporate the organic solvent. The resulting particles were analyzed by DLS, without the centrifugation and cleaning procedure utilized for later methods.

Nanoparticle production by nanoprecipitation

Nanoparticles prepared by nanoprecipitation followed a similar procedure as that described in Chapter 5, but on a smaller scale and with additional cleaning steps (Chang et al., 2009; Cheng et al., 2007). PHODN3 was dissolved in one of several organic solvent systems: tetrahydrofuran

(THF), dioxane, acetone, 15% dimethyl sulfoxide (DMSO) in acetone, 15% ethanol in acetone, and dimethylformamide. Polymer solutions were added dropwise to rapidly stirring SDS solutions

(4.44 mg mL-1) in one of two ratios; for a 1:1 ratio, 2 mL of 1 mg mL-1 polymer solution was added

206 to 2 mL SDS solution, and for a 1:10 ratio 1 mL of 1 mg mL-1 polymer solution was added to 10 mL SDS. The combined solutions were stirred for 24 hours to evaporate the organic solvents.

To remove the SDS from the nanoparticles, 1 mL of nanoparticle solution was transferred to a 1.5 mL microcentrifuge tube and centrifuged for 30 min at 16,100 × g (Eppendorf 5415R).

The supernatant was carefully removed by aspiration and analyzed by DLS, while the pelleted nanoparticles were resuspended in 1 mL Nanopure water and centrifuged again. This supernatant was discarded, and the pellets resuspended in 1 mL water for DLS analysis. For 1:10 samples, 4 aliquots were centrifuged, resuspended in 0.25 mL, and combined to obtain more nanoparticles.

Both the initial supernatant and the cleaned nanoparticle solutions were filtered through 0.2 um

Nylon filters prior to DLS.

References

Budhian, A., Siegel, S.J., Winey, K.I., 2007. Haloperidol-loaded PLGA nanoparticles: Systematic

study of particle size and drug content. International Journal of Pharmaceutics 336, 367–

375. https://doi.org/10.1016/j.ijpharm.2006.11.061

Chang, J., Jallouli, Y., Kroubi, M., Yuan, X., Feng, W., Kang, C., Pu, P., Betbeder, D., 2009.

Characterization of endocytosis of transferrin-coated PLGA nanoparticles by the blood–

brain barrier. International Journal of Pharmaceutics, Challenges for Nanotechnology in

Delivery Imaging 379, 285–292. https://doi.org/10.1016/j.ijpharm.2009.04.035

Cheng, J., Teply, B.A., Sherifi, I., Sung, J., Luther, G., Gu, F.X., Levy-Nissenbaum, E., Radovic-

Moreno, A.F., Langer, R., Farokhzad, O.C., 2007. Formulation of Functionalized PLGA-

PEG Nanoparticles for In Vivo Targeted Drug Delivery. Biomaterials 28, 869–876.

https://doi.org/10.1016/j.biomaterials.2006.09.047

207

Mainardes, R.M., Evangelista, R.C., 2005. PLGA nanoparticles containing praziquantel: effect of

formulation variables on size distribution. Int J Pharm 290, 137–144.

https://doi.org/10.1016/j.ijpharm.2004.11.027

208

Appendix C: Gel permeation chromatography data

The weight average (Mw) and number average (Mn) molecular weights for each sample were determined by gel permeation chromatography (GPC) as described previously (Pinto et al., 2016).

Briefly, PHA solutions of approximately 1.0 g L-1 were prepared by dissolution in chloroform and passed through a syringe filter (0.45 μm PTFE). Samples were injected (50 μL) into a Shimadzu

LC-20AD liquid chromatograph equipped with a Shimadzu SIL-20A autosampler, a Shimadzu

CTO-20A column oven, and a Shimadzu RID-10A refractive index detector. Samples were passed through an 8 x 50 mm styrenedivinylbenzene (SDV) guard column (5 μm particles; Polymer

Standards Service) and an 8 x 300 mm SDV analytical column (5 μm particles; mixed bed porosity; max molecular weight 1E6 Da; Polymer Standards Service product sda0830051lim). The column oven was maintained at 40°C with a 1 mL min-1 mobile phase of chloroform. Molecular weight standards of polystyrene with a narrow polydispersity index were used for calibration. Shimadzu’s

LCsolution software was used to analyze the data, and these sample chromatograms were generated using RStudio. Figs. C1-C7 show data generated from experiments in Chapter 2, Figs.

C8-C19 were generated from experiments in Chapter 4.

Fig. C1: GPC chromatogram of purified poly(3-hydroxybutyrate) (C4) produced by RSC02, plotted as normalized intensity vs retention time. No data available for LSBJ.

209

A B

Fig. C2: GPC chromatogram of purified poly(3-hydroxyvalerate) (C5) produced by LSBJ (A) and RSC02 (B), plotted as normalized intensity vs retention time.

A B

Fig. C3: GPC chromatogram of purified poly(3-hydroxyhexanoate) (C6) produced by LSBJ (A) and RSC02 (B), plotted as normalized intensity vs retention time.

210

A B

Fig. C4: GPC chromatogram of purified poly(3-hydroxyheptanoate) (C7) produced by LSBJ (A) and RSC02 (B), plotted as normalized intensity vs retention time.

A B

Fig. C5: GPC chromatogram of purified poly(3-hydroxyoctanoate) (C8) produced by LSBJ (A) and RSC02 (B), plotted as normalized intensity vs retention time.

211

A B

Fig. C6: GPC chromatogram of purified poly(3-hydroxydecanoate) (C10) produced by LSBJ (A) and RSC02 (B), plotted as normalized intensity vs retention time.

A B

Fig. C7: GPC chromatogram of purified poly(3-hydroxydodecanoate) (C12) produced by LSBJ (A) and RSC02 (B), plotted as normalized intensity vs retention time.

212

Fig. C8: GPC chromatogram of purified poly(3-hydroxydecanoate) produced by LSBJ via fed- batch fermentation (Trial 9A). Data plotted as normalized intensity vs retention time.

Fig. C9: GPC chromatogram of purified poly(3-hydroxydecanoate) produced by RSC02 via fed- batch fermentation (Trial 10B). Data plotted as normalized intensity vs retention time.

213

Fig. C10: GPC chromatogram of purified poly(3-hydroxydecanoate) produced by LSBJ via fed- batch fermentation (Trial 12A). Data plotted as normalized intensity vs retention time.

Fig. C11: GPC chromatogram of purified poly(3-hydroxydecanoate) produced by LSBJ via fed- batch fermentation (Trial 13A). Data plotted as normalized intensity vs retention time.

214

Fig. C12: GPC chromatogram of purified poly(3-hydroxydecanoate) produced by LSBJ via fed- batch fermentation (Trial 15A). Data plotted as normalized intensity vs retention time.

Fig. C13: GPC chromatogram of purified poly(3-hydroxydecanoate) produced by LSBJ via fed- batch fermentation (Trial 16A). Data plotted as normalized intensity vs retention time.

215

Fig. C14: GPC chromatogram of purified poly(3-hydroxydecanoate) produced by LSBJ via fed- batch fermentation (Trial 17A). Data plotted as normalized intensity vs retention time.

Fig. C15: GPC chromatogram of purified poly(3-hydroxydecanoate) produced by LSBJ via fed- batch fermentation (Trial 18A). Data plotted as normalized intensity vs retention time.

216

Fig. C16: GPC chromatogram of purified poly(3-hydroxyhexanoate) produced by LSBJ via fed- batch fermentation (Trial 19A). Data plotted as normalized intensity vs retention time.

Fig. C17: GPC chromatogram of purified poly(3-hydroxyoctanoate) produced by LSBJ via fed- batch fermentation (Trial 20A). Data plotted as normalized intensity vs retention time.

217

Fig. C18: GPC chromatogram of purified poly(3-hydroxy-10-undecenoate) produced by LSBJ via fed-batch fermentation (Trial 21A). Data plotted as normalized intensity vs retention time.

Fig. C19: GPC chromatogram of purified poly(3-hydroxy-10-bromodecanoate) produced by LSBJ via fed-batch fermentation (Trial 22A). Data plotted as normalized intensity vs retention time.

218

References

Pinto, A., Ciesla, J.H., Palucci, A., Sutliff, B.P., Nomura, C.T., 2016. Chemically Intractable No

More: In Vivo Incorporation of “Click”-Ready Fatty Acids into Poly-[(R)-3-

hydroxyalkanoates] in Escherichia coli. ACS Macro Letters 5, 215–219.

https://doi.org/10.1021/acsmacrolett.5b00823

219

Appendix D: Bioreactor Run Parameters

The representative data presented in this appendix are from the dissolved oxygen (DO), pH, temperature (T), agitation speed, and air sparging rate which were monitored and recorded every

30 seconds during each fermentation. One replicate per Trial (starting with 9) is shown.

Fig. D1: Trial 9A. Bioreactor fermentation of E. coli LSBJ harboring pBBRC1J4SII, with 6 g decanoic acid added at 0.22 g h-1.

220

Fig. D2: Trial 9B. Bioreactor fermentation of E. coli LSBJ harboring pBBRC1J4SII, with 6 g decanoic acid added at 0.22 g h-1.

221

Fig. D3: Trial 10A. Bioreactor fermentation of E. coli RSC02 harboring pBBRC1J4SII, with 6 g decanoic acid added at 0.22 g h-1.

222

Fig. D4: Trial 11A. Bioreactor fermentation of E. coli LSBJ CRP* harboring pBBRC1J4SII, with 6 g decanoic acid added at 0.22 g h-1.

223

Fig. D5: Trial 12A. Bioreactor fermentation of E. coli LSBJ harboring pBBRC1J4SII, with 6 g decanoic acid added at 0.22 g h-1. Rate of glucose feed slowed to 9.7 mL h-1 during phase 2.

224

Fig. D6: Trial 13A. Bioreactor fermentation of E. coli LSBJ harboring pBBRC1J4SII, with 6 g decanoic acid added at 0.22 g h-1. Media formulation changed to that described in Chapter 4, high-density fed-batch production. Rate of glucose feed slowed to 9.7 mL h-1 during phase 2. First run where culture survived for the entire 48 h (O2 demand slowly increased between 13 and 48 h).

225

Fig. D7: Trial 14. Bioreactor fermentation of E. coli LSBJ harboring pBBRC1J4SII, with 12 g decanoic acid added at 0.44 g h-1. Media described in Chapter 4, high-density fed-batch production. Rate of glucose feed slowed to 9.7 mL h-1 during phase 2.

226

Fig. D8: Trial 15A. Bioreactor fermentation of E. coli LSBJ harboring pBBRC1J4SII, with 12 g decanoic acid added at 0.44 g h-1. Media described in Chapter 4, high-density fed-batch production, with 3.0 g of yeast extract. Rate of glucose feed slowed to 9.7 mL h-1 during phase 2.

227

Fig. D9: Trial 16A. Bioreactor fermentation of E. coli LSBJ harboring pBBRC1J4SII, with 15 g decanoic acid added at 0.55 g h-1. Media described in Chapter 4, high-density fed-batch production, with 3.0 g of yeast extract. Rate of glucose feed slowed to 9.7 mL h-1 during phase 2.

228

Fig. D10: Trial 17A. Bioreactor fermentation of E. coli LSBJ harboring pBBRC1J4SII, with 20 g decanoic acid added at 0.65 g h-1. Media described in Chapter 4, high-density fed-batch production, with 3.0 g of yeast extract. Initial rate of glucose feed reduced to 12.9 mL h-1, slowed to 8.7 mL h-1 during phase 2.

229

Fig. D11: Trial 18. Bioreactor fermentation of E. coli LSBJ harboring pBBRC1J4SII, with 25 g decanoic acid added at 0.75 g h-1. Media described in Chapter 4, high-density fed-batch production, with 3.0 g of yeast extract. Initial rate of glucose feed reduced to 11.4 mL h-1, slowed to 7.7 mL h-1 during phase 2.

230

Fig. D12: Trial 19. Bioreactor fermentation of E. coli LSBJ harboring pBBRC1J4SII, with 20 g hexanoic acid added at 0.65 g h-1. Media described in Chapter 4, high-density fed-batch production, with 3.0 g of yeast extract. Initial rate of glucose feed reduced to 12.9 mL h-1, slowed to 8.7 mL h-1 during phase 2.

231

Fig. D13: Trial 20. Bioreactor fermentation of E. coli LSBJ harboring pBBRC1J4SII, with 20 g octanoic acid added at 0.65 g h-1. Media described in Chapter 4, high-density fed-batch production, with 3.0 g of yeast extract. Initial rate of glucose feed reduced to 12.9 mL h-1, slowed to 8.7 mL h-1 during phase 2.

232

Fig. D14: Trial 21. Bioreactor fermentation of E. coli LSBJ harboring pBBRC1J4SII, with 20 g 10-undecenoic acid added at 0.65 g h-1. Media described in Chapter 4, high-density fed-batch production, with 3.0 g of yeast extract. Initial rate of glucose feed reduced to 12.9 mL h-1, slowed to 8.7 mL h-1 during phase 2.

233

Fig. D15: Trial 22. Bioreactor fermentation of E. coli LSBJ harboring pBBRC1J4SII, with 20 g 10-bromodecanoic acid added at 0.65 g h-1. Media described in Chapter 4, high-density fed-batch production, with 3.0 g of yeast extract. Initial rate of glucose feed reduced to 12.9 mL h-1, slowed to 8.7 mL h-1 during phase 2.

234

Fig. D16: Trial 23. Bioreactor fermentation of E. coli LSBJ harboring pBBRC1J4SII, with 15.4 g octanoic acid and 4 g decanoic acid added at 0.632 total g h-1. Media described in Chapter 4, high- density fed-batch production, with 3.0 g of yeast extract. Initial rate of glucose feed reduced to 12.9 mL h-1, slowed to 8.7 mL h-1 during phase 2.

235

Fig. D17: Trial 24. Bioreactor fermentation of E. coli LSBJ harboring pBBRC1J4SII, with 15.4 g octanoic acid and 4.95 g 10-azidodecanoic acid added at 0.663 total g h-1. Media described in Chapter 4, high-density fed-batch production, with 3.0 g of yeast extract. Initial rate of glucose feed reduced to 12.9 mL h-1, slowed to 8.7 mL h-1 during phase 2.

236

Appendix E: HPLC Sample Chromatograms

The data presented in this appendix are a sample high performance liquid chromatography (HPLC) chromatogram of paclitaxel and the standard curve generated for paclitaxel quantification. HPLC was performed as described in Chapter 5. Paclitaxel absorbance was measured at 227 nm.

1.20E+07

1.00E+07 )

227 8.00E+06

y = 1,985,106.5255x 6.00E+06 R² = 0.9999

4.00E+06 Peak Integration (Abs Integration Peak

2.00E+06

0.00E+00 0.000 1.000 2.000 3.000 4.000 5.000 Mass of Taxol (μg)

Fig. E1: Calibration curve generated for the quantification of paclitaxel.

237 mV

Fig. E2: Sample HPLC chromatogram of paclitaxel encapsulated in PHODN3 nanoparticle (dissolved in acetonitrile). Photocopied from lab notebook due to campus shutdown (COVID-19).

238

Appendix F: Additional Supplemental Data

This section contains additional supplemental data published with Scheel et al. 2016 (Chapter 2), as well as GC chromatograms for poly[(R)-3-hydroxy-10-bromodecanoate] (PHDBr) and a schematic of the bioreactor used for experiments in Chapter 4.

Table F1: Poly(3-hydroxydecanoate) yield comparison between LSBJ, RSC02, RSC04, and RSC06 Strain CDWa (g L-1) PHAa (wt%) PHA Concentrationa (mg L-1) LSBJ 1.13 ± 0.02 17.4 ± 3.7 196 ± 40 RSC02 1.33 ± 0.04* 26.4 ± 4.7* 353 ± 77* RSC04 0.93 ± 0.00* 16.5 ± 1.9 154 ± 18 RSC06 0.99 ± 0.02* 17.9 ± 3.0 179 ± 31 a All values are averages of biological triplicate experiments plus or minus the standard deviation about those averages. * Denotes statistically significant difference compared to LSBJ (Student’s t-test, two-tailed, α = 0.05).

uV (x10,000) 5.0 Chromatogram

4.0

3.0

2.0

1.0

0.0

5.0 10.0 15.0 20.0 25.0 30.0 min

Fig. F1: Sample GC chromatogram of purified PHDBr subjected to methanolysis as described in Chapter 4. Several peaks were generated by the polymer, possibly due to side reactions from methanolysis. Broad peak at ~28 – 30 min is likely the 3-hydroxy-10-bromodecanoate methyl ester; however, this region is often populated by background peaks from methanolyzed cell components and the peak at 25 min was calibrated as a proxy for the polymer.

239

uV (x10,000) 5.0 Chromatogram

4.0

3.0

2.0

1.0

0.0

5.0 10.0 15.0 20.0 25.0 30.0 min

Fig. F2: Sample GC chromatogram of control lyophilized cells (polymer-free, 12h fermentation) subjected to methanolysis. Note 2 peaks at ~28.5 min; these background peaks grow with increased fermentation time and are the reason for choosing the peak at 25 min for PHDBr calibration.

uV (x10,000) 5.0 Chromatogram

4.0

3.0

2.0

1.0

0.0

5.0 10.0 15.0 20.0 25.0 30.0 min

Fig. F3: Sample GC chromatogram of methanolyzed cells from Trial 22 at 48h (PHDBr production, Table 4.2).

240

Conc.(x10) 1.00

0.75

0.50 Y = aX + b a = 1.574496e-004 0.25 b = 0.0 R^2 = 0.9918222 0.00 R = 0.9959027 0 10000 20000 30000 40000 50000 Area Fig. F4: Calibration curve generated for quantification of methanolyzed PHDBr.

Fig. F5: Schematic of the New Brunswick BioFlo 310 2L bioreactor vessel used in Chapter 4. Four sensors were installed to measure dissolved oxygen (DO), pH, temperature (Temp), and the culture level (for foam detection). Glucose and fatty acid feeds, as well as concentrated ammonium hydroxide and 10% anti-foam B in water, were added by peristaltic pumps as described in Chapter 4.

241

Appendix G: Peer-reviewed publication covers

242

243

244

245

246

247

248

Ryan A. Scheel – Curriculum Vitae State University of New York College of Environmental Science and Forestry (SUNY ESF) 1 Forestry Drive ● Syracuse, NY, 13210 ● (585) 406-4467 [email protected]

EDUCATION 2015-2020 SUNY ESF Syracuse, NY Advisor: Dr. Christopher T. Nomura Ph.D. Department of Chemistry, concentration in Biochemistry. Research focused on bacterial gene regulation for (1) membrane transport in Pseudomonas aeruginosa and (2) biosynthesis of polyhydroxyalkanoates in Escherichia coli.

2011-2014 SUNY ESF Syracuse, NY Bachelor of Science in Environmental Biology with Honors, minoring in Chemistry and Biotechnology, Magna Cum Laude (GPA 3.58). Participated in honors research culminating in a thesis and presentation. Thesis: Deregulation of fatty acid transport in Escherichia coli for enhanced control of biodegradable plastic copolymer production.

2007-2011 SUNY Monroe Community College Rochester, NY Associates in Liberal Arts and Sciences (GPA 3.97)

PEER-REVIEWED PUBLICATIONS

1. Scheel RA, Fusi AD, Min BC, Thomas CM, Ramarao BV, Nomura CT. Increased production of the value-added biopolymers poly(R-3-hydroxyalkanoate) and poly(γ-glutamic acid) from hydrolyzed paper recycling waste fines. Front. Bioeng. Biotechnol. 7 (2019). 2. Salamanca-Cardona L, Scheel RA, Mizuno K, Bergey NS, Stipanovic AJ, Matsumoto K, Taguchi S, Nomura CT. Effect of acetate as a co-feedstock on the production of poly(lactate-co-3-hydroxyalkanoate) by pflA- deficient Escherichia coli RSC10. J. Biosci. Bioeng. 123: 547-554 (2017). 3. Salamanca-Cardona L, Scheel RA, Bergey NS, Stipanovic AJ, Matsumoto K, Taguchi S, Nomura CT. Consolidated bioprocessing of poly(lactate-co-3-hydroxybutyrate) from xylan as a sole feedstock by genetically-engineered Escherichia coli. J. Biosci. Bioeng. 122: 406-414 (2016). 4. Scheel RA, Ji L, Lundgren BR, Nomura CT. Enhancing poly(3-hydroxyalkanoate) production in Escherichia coli by the removal of the regulatory gene arcA. AMB Express 6 (2016). 5. Chen Q, Yang B, Wang H, He F, Gao Y, Scheel RA. Soil microbial community toxic response to atrazine and its residues under atrazine and lead contamination. Environ. Sci. Pollut. Res. 22: 996–1007 (2015). 6. Lundgren BR, Harris JR, Sarwar Z, Scheel RA, Nomura CT. The metabolism of (R)-3-hydroxybutyrate is regulated by the enhancer-binding protein PA2005 and the alternative sigma factor RpoN in Pseudomonas aeruginosa PA01. Microbiology 161: 2232-2242 (2015). 7. Salamanca-Cardona L, Scheel RA, Lundgren BR, Stipanovic AJ, Matsumoto K, Taguchi S, Nomura CT. Deletion of the pflA gene In Escherichia coli LS5218 and its effects on the production of polyhydroxyalkanoates using beechwood xylan as a feedstock. Bioengineered 5: 284-287 (2014).

RESEARCH EXPERIENCE 2015-Present Doctoral Candidate, Department of Chemistry SUNY ESF Advisor: Dr. Christopher T. Nomura Research focus: Metabolic engineering of Escherichia coli to enhance biosynthesis of biodegradable poly(R-3-hydroxyalkanoates) (PHA), and characterization of a gene regulatory network in Pseudomonas aeruginosa. Currently focused on optimizing high-density fermentation methods for medium chain length (MCL) PHA biosynthesis, synthesis of functionalized PHA nanoparticles for targeted drug delivery,

249

and engineering a novel metabolic pathway for poly(4-hydroxybutyrate)-co-MCL PHA for high-density fermentation.

Summer 2014 Research Internship, Third Millennium Alliance Manabi, Ecuador Advisor: Jordan Trujillo Pursued an ethnobotanical research project on natural insect repellents in the coastal rainforest of Ecuador, worked closely with local community members, and presented research results in Spanish to the village of Camarones. Volunteered for the CEIBA foundation assisting Camarones women with piñon planting.

2012-2014 Undergraduate Research, Nomura Research Group SUNY ESF Advisors: Dr. Christopher T. Nomura, Dr. Benjamin R. Lundgren, Dr. Ryan Tappel Research focus: Genetic engineering of Escherichia coli to enhance uptake and metabolism of related and un-related substrates for poly(R-3-hydroxyalkanoate) biosynthesis. Resulted in preliminary data for graduate thesis research.

FELLOWSHIPS AND SCHOLARSHIPS 2016-2018 NIH Predoctoral Fellowship (F31) SUNY ESF Awarded for research proposal investigating a putative uncharacterized enhancer binding protein and its regulon in Pseudomonas aeruginosa. 2011-2014 Phi Theta Kappa Scholarship, International Honor Society for 2- and 4-Year Colleges Awarded for exemplary academic achievements.

HONORS AND AWARDS May 2019 Excellence in Academic Research Award, ESF Graduate Student Association SUNY ESF May 2017 2nd Place Research Presentation, 9th NYS Biotechnology Symposium Fall 2015 1st Place Eastman Chemical Company Graduate Research Competition 2013-2014 Upper Division Honors Program 2011-2014 Dean’s List (Fall semesters 2011-2013, Spring of 2014)

TEACHING EXPERIENCE Spring 2018 Guest Lecturer, Biochemistry SUNY ESF Prepared and presented a lecture on fatty acid biosynthesis and an introduction to polyketide biosynthesis for senior/graduate level course.

Fall 2017 Lab Coordinator/Teaching Assistant, Biochemistry SUNY ESF Maintained a lab inventory and ordered necessary chemicals for experiments, prepared buffers and reagents for students, set up research instruments and supervised their use, presented pre-lab lectures on enzyme characterization and kinetics, and instructed a 10-12 hour/week lab course.

Spring 2015 Teaching Assistant, General Chemistry SUNY ESF Supervised and instructed 2 lab sections in general chemistry techniques. Presented pre-lab lectures on principles of chemistry, assisted students in the lab, and mentored students during weekly office hours.

MENTORING EXPERIENCE 2014-Present Undergraduate Researcher Mentees SUNY ESF Truong Ho (BS Biochemistry, 2021), Yuki Kageyama (MS Biological Chemistry and Engineering, International Research Experience Intern Fall 2019), Jennifer Callan (High School Mentorship Program,

250

Summer 2019), Alex Kirschner (BS Biotechnology, 2019), Jessica Masisak (BS Biochemistry, 2019), Michael Glinski (BS Bioprocess Engineering, 2018), Alexander Fusi (Biotechnology, 2018), Dylan Rebolini (BS Biotechnology, 2017), Joel Wallace (REU Student, Summer 2017), James Capanegro (Biotechnology, 2016), Ellen Conti (BS Bioprocess Engineering, 2016), Liyuan Ji (SUNY Stem Passport Intern, Summer 2016), Seamus McKenney (BS Biotechnology, 2016).

LEADERSHIP EXPERIENCE 2017-2019 President, Graduate Student Association (GSA) SUNY ESF Organized and chaired biweekly senate meetings in accordance with the GSA Bylaws and Robert’s Rules of Order. Prepared and circulated meeting agendas and was responsible for email communication to the graduate student body. Served as voting representative to the ESF Leadership Council and Academic Governance. Led the initiative to transform the organization from a club to an independent 501(3)(c) non-profit. Represented the GSA at biannual meetings of the SUNY Student Association, the university- wide student governance body of SUNY’s 64 campuses.

2016-2017 Representative to Undergraduate Student Association, GSA SUNY ESF Acted as liaison between the Graduate and Undergraduate Student Associations. Served as one of eight voting members for GSA at Academic Governance and served on the Academic Governance committee on Instructional Quality and Academic Standards. Authored a resolution on Graduate Student Safety and planned social events on and off campus as a member of the GSA Social Activities Committee.

PROFESSIONAL EXPERIENCE 2007-2017 Food Service Experience, 10 years Rochester & Syracuse, NY Worked in both front- and back-of-house positions at various restaurants while enrolled in college.

PRESENTATIONS August 2018 25th Bio-Environmental Polymer Society Meeting, Invited Speaker Rensselaer Polytechnic “Engineering Escherichia coli for enhanced biosynthesis of poly[(R)-3-hydroxyalkanoates]”

August 2017 20th CBE Graduate Research Symposium, Poster presentation University of Buffalo “Enhancing poly(3-hydroxyalkanoate) production in Escherichia coli by the removal of the regulatory gene arcA”

May 2017 9th NYS Biotechnology Symposium, 2nd Place, Poster presentation SUNY ESF “Enhancing poly(3-hydroxyalkanoate) production in Escherichia coli by the removal of the regulatory gene arcA”

May 2016 8th NYS Biotechnology Symposium, Poster presentation SUNY ESF “Metabolic engineering of Escherichia coli LSBJ for improved production of poly (3-hydroxyalkanoates)”

March 2016 Stevenson Biomaterials Lecture Series, Poster presentation Syracuse University “Metabolic engineering of Escherichia coli LSBJ for improved production of poly (3-hydroxyalkanoates)”

June 2015 Boston Bacterial Meeting, Poster presentation Harvard University “Metabolic engineering of Escherichia coli LSBJ for improved production of poly (3-hydroxyalkanoates)”

April 2015 Spotlight on Research, Graduate Poster presentation SUNY ESF “Metabolic engineering of Escherichia coli LSBJ for improved production of poly (3-hydroxyalkanoates)” April 2014 Spotlight on Research, Undergraduate Poster presentation SUNY ESF

251

“Deregulation of fatty acid transport in Escherichia coli for enhanced control of biodegradable plastic copolymer production”

LANGUAGES American English – Native/Bilingual (ILR Level 5) Spanish – Limited working proficiency (ILR Level 2+)

REFERENCES Christopher T. Nomura, Ph.D. (he/him/his) Vice President for Research Professor of Biochemistry | SUNY College of Environmental Science and Forestry 1 Forestry Drive, Syracuse, NY 13210 Phone: (315) 470-6606 Email: [email protected]

Atahualpa Pinto, Ph.D. (he/him/his) Visiting Professor of Chemistry | Bard College 30 Campus Road, Annandale-on-Hudson, NY 12504 Phone: (845) 752-2311 Fax: (845) 752-2381 Email: [email protected]

Zaara Sarwar, Ph.D. (she/her/hers) Assistant Professor of Biology | The College of New Jersey 2000 Pennington Rd, Ewing, NJ 08628 Phone: (609) 771-2457 Email: [email protected]

252