ABSTRACT Title of Document: CONTEXTUALIZATION OF THE E. COLI LSR SYSTEM: RELATIVE ORTHOLOGY, RELATIVE QS ACTIVITY, AND EMERGENT BEHAVIOR David Nathan Quan, Doctor of Philosophy, 2014 Directed By: Professor William E. Bentley Fischell Department of Bioengineering Within bacterial consortia there exist innumerable combinatorial circumstances, some of which may tip the scale toward pathogenicity, some of which may favor asymptomatic phenotypes. Indeed, the lines and intersections between commensal, pathogenic, and opportunistic bacteria are not always clean. As a foothold to mediate pathogenicity arising from consortia, many have puzzled at communication between bacteria. Primary among such considerations is quorum sensing (QS). Analogous to autocrine signaling in multicellular organisms, QS is a self-signaling process involving small molecules. Generally, QS activation is believed to have pleiotropic effects, and has been associated with numerous pathogenic phenotypes. The research herein focuses on autoinducer-2 (AI-2) based QS signaling transduced through the Lsr system. Produced by over 80 species of bacteria, AI-2 is believed to be an interspecies signaling molecule. Outside of the marine bacteria genera Vibrio and Marinomonas, the only known AI-2 based QS transduction pathway is the Lsr system. We sought to deepen the characterization of the Lsr system in contexts outside of the batch cultures in which it was originally defined. First, we interrogated E. coli K-12 W3110 Lsr system orthologs relative to the same strain’s lac system. Both systems are induced by the molecule which they import and catabolize. We searched for homologs by focusing on the gene order along a genome, as gene arrangement can bear signaling consequences for autoregulatory circuits. We found that the Lsr system signal was phylogenetically dispersed if not particularly deep, especially outside of Enterobacteriales and Pasteurellaceaes, indicating that the system has generally been conferred horizontally. This contrasts with the lac system, whose signal is strong but limited to a select group of highly related enterobacteria. We then modeled the Lsr system with ODEs, revealing bimodality in silico, bolstering preliminary experimental evidence. This bifurcated expression was seen to depend upon nongenetic heterogeneity, which we modeled as a variation of a single compound parameter, basal, representing the basal rate of AI-2 flux into the cell through a low flux pathway. Moreover, in our finite difference-agent based models, bimodal expression could not arise from spatial stochasticity alone. This lies in contrast with the canonical LuxIR QS system, which employs an intercellular positive feedback loop to activate the entire population. We examined the consequences of this contrast, by modeling both systems under conditions of colony growth using finite difference-agent based methods. We additionally investigated the confluence of Lsr signaling with chemotactic sensitivity to AI-2, which has been demonstrated in E. coli. Finally, the consequences of bimodality in interspecies interactions were assessed by posing two populations containing different Lsr systems against each other. While few natural consortia consist of only two interacting bacteria, these studies indicate that AI-2 based Lsr signaling may mediate a multitude of transitional intraspecies and interspecies bacterial dynamics, the specifics of which will vary with the context and the homologs involved. CONTEXTUALIZATION OF THE E. COLI LSR SYSTEM: RELATIVE ORTHOLOGY, RELATIVE QS ACTIVITY, AND EMERGENT BEHAVIOR By David Nathan Quan Dissertation submitted to the Faculty of the Graduate School of the University of Maryland, College Park, in partial fulfillment of the requirements for the degree of Doctor of Philosophy 2014 Advisory Committee: Professor William E. Bentley, Chair Associate Professor Michael P. Cummings Associate Professor Nam Sun Wang Associate Professor Adam Hsieh Assistant Professor Ganesh Sriram © Copyright by David Nathan Quan 2015 ii Dedication As a token of love and appreciation, to my parents, to my brother, and my broader family who have sacrificed so much, both small and great—to their unrelenting and unconditional support. iii Acknowledgements I wish to acknowledge all the members of my dissertation committee for their time, advice, and commitment to myself, their own graduate students, and the university. I especially recognize my committee chair and advisor, Dr. William Bentley, under whose guidance I have been allowed to follow my nose and in whose lab I have enjoyed a stimulating research environment. I also wish to acknowledge my fellow lab members both present and past, whose work serves as fundamental bedrock for my own, and whose ideas I have come to recount as readily as my own. In particular, I wish to acknowledge Dr. Chen-Yu Tsao whose development of pCT5 and pCT6 has been utterly important for our lab and has resulted in compelling basic science. I also acknowledge the work of those in the quorum sensing field as whole—it is only by standing upon their shoulders that I can view any forest at all. I also wish to gratefully acknowledge my mentors at the Center for Cell Dynamics, the members of the Alliance for Cellular Signaling, and Dr. Gary Jarvis for giving me a wide berth when trying to do science during my formative years. In particular, I would like to recognize a former mentor at the Center for Cell Dynamics, Dr. Jonathon Alberts, whose Sim2D platform was modified and heavily relied upon here. Finally, I acknowledge the reader: I hope you find what you need or perhaps even something interesting or worthwhile. iv Table of Contents Dedication ......................................................................................................................... ii Acknowledgements............................................................................................................iii Table of Contents................................................................................................................iv List of Tables .....................................................................................................................vi List of Figures .................................................................................................................. vii Chapter 1: Introduction ..................................................................................................... 1 1.1 Background.......................................................................................................... 1 1.1.1 QS in Vibrio Harveyi….………………………………….……………... 5 1.1.2 LuxIR QS…….….………………………………………………………. 7 1.1.3 AI-2 QS………….…………………………………………………….. 10 1.1.4 Lsr Regulon.……………………………………………………………. 15 1.1.5 Known Lsr Associated Phenotypes.……………………………………. 16 1.1.6 Broader AI-2 Associated Phenotypes…………...….………………….. 17 1.1.7 QS in an Ecological Context….……………………………………..…. 20 1.2 Research Motivation……………………………………………………………22 1.3 Global Objective, Global Hypothesis, and Specific Aim……………………... 26 1.4 Dissertation Outline…………………………………………………………… 27 Chapter 2: Comparison of Homolog Identification for the Sugar Importing lac System and the QS Lsr System from E. coli K-12 W3110…………………………………….......... 28 2.1 Abstract……………………………………………………………………….. 28 2.2 Introduction…………………………………………………………………… 30 2.3 Methods and Algorithm...................................................................................... 34 2.3.1 Input……………………………………………………………………. 37 2.3.2 Scoring Heuristic………………………………………………………. 37 2.3.3 Weak and Stringent Criteria…………………………………………… 38 2.3.4 Ancillary LMNAST Search Tools……………………………………... 38 2.4 Results……………….........................................................................................40 2.4.1 E. coli K-12 W3110 lac Operon Query………………………………… 40 2.4.2 E. coli K-12 W3110 Lsr System Query………………………………… 48 2.4.3 Analysis of Lsr System Search Results……………...……………….... 57 2.4.3.1 Putative Lsr System in Rhizobiales…………………………….. 59 2.5 Concluding Remarks...………………………………………………………… 60 2.6 Supplemental Material………………………………………………………… 64 Chapter 3: Quorum desynchronization leads to bimodality and patterned behaviors in microbial consortia …...................................................................................................... 74 3.1 Abstract.............................................................................................................. 74 3.2 Introduction........................................................................................................ 75 3.3 Methods.............................................................................................................. 79 3.3.1 Modeled Cell Behaviors……..……………………………………..........79 3.5.1.1 Chemotactic Swimming……………………………………….....79 3.5.1.2 Colony Growth…………………………………………………...79 3.5.1.3 LuxIR/AHL QS………………………………………………..…80 3.5.1.4 Lsr/AI-QS……………………………………………………...…80 3.3.2 Simulation Variants: Gene Deletions and Mixed Populations…………..80 v 3.4 Results................................................................................................................ 82 3.4.1 Lsr Autoinduction in Pure Cultures……………………………….……..82 3.4.2 Pattern formation in QS systems: LuxIR vs Lsr……….………………...84 3.4.3 Cell Motility – Lsr QS based pattern emergence………………...………85 3.4.4 Mixed Population Simulations…………………………………………...87 3.5 Discussion………………………………………………………………………91 3.6 Concluding Remarks……………………………………………………………96 3.7 Supplemental Information………………………………………………………98 3.7.1 Text and Discussion ……………………………….……………………98 3.7.2 Methods……………………...…………………………………………..99 3.7.2.1 ODE Model…...……….…………………………………………99 3.7.2.1.1 Generalities and Scope.….…….…………………………...99 3.7.2.1.2 mRNA Expression……………………………..………….100 3.7.2.1.3 Protein Synthesis……………………………….…………101 3.7.2.1.4 mRNA and Protein Degredation...……………………..…101 3.7.2.1.5 Cell Growth………………………………………….....…101 3.7.2.1.6 AI-2 Transport......…………………………………..…….102 3.7.2.1.7 AI-2 Degredation and Synthesis…………..………………103 3.7.2.1.8 Equations…………………………………..……………...103 3.7.2.1.9 Parameter Values.....………………………………..……..105 3.7.2.1.10 Initial Values………………………………………..……107 3.7.2.1.11 Numerical Solution…………………………………..…..108 3.7.2.2 Finite Difference-Agent Based Model…………………………...108 3.7.2.2.1 Modeled Environment………..…………………………...108 3.7.2.2.2 Adaptation of Equations and Solutions……………..…….108 3.7.2.2.3 Cell ODE Numerical Solution Method…...………..……. 109 3.7.2.2.4 Cell Division……………………………………...……….109 3.7.2.2.5 Diffusion……………………………………………..……110 3.7.2.2.6 Time Interval Order………..………………………....….. 110 3.7.3 Results…….…………………………..………………………………111 3.7.3.1 Numerical solutions to ODE.………………………………….111 3.7.3.2 System Sensitivity to basal rate of AI-2 uptake………………..111 3.7.3.3 Two sets of Lsr with different rates of basal AI-2 uptake……..115 3.7.3.4 Full population of cells with Lsr in a finite difference environment …………………………………………………………………………117 3.7.3.5 Minimal role of spatial heterogeneity of AI-2……………...…119 3.7.3.6 Agreement between numerical ODE and finite difference agent based solutions………………………………………………………...121 3.7.3.7 Heterogeneity of local Lsr and LuxIR QS activation in growing colonies………………………………………….………….…………121 3.7.3.8 Evaluation of clustering when Lsr QS is coupled to AI-2 chemoattraction…………………….……………………………..…..125 3.7.3.9 Motility mode feedback onto population activation as a function of cell-cell distance…………………………………………………...126 vi Chapter 4: Conclusions................................................................................................... 130 References...................................................................................................................... 133 vii List of Tables Table 2-S1. Accompaniment for Figure 2-S1, Trackback plots for Lsr system LMNAST extended window stringent search hits. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65 Table 2-S2. Results from three LMNAST searches for Lsr system homologs. . . . . 69 Table 3-S1. Estimated parameter values. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106 viii List of Figures Figure 1-1. Synthesis of AHL and DPD. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3 Figure 1-2. QS molecule activation of QS signaling pathway. . . . . . . . . . . . . . . . . . . 4 Figure 1-3. LuxIR activity at single cell and multicellular hierarchies. . . . . . . . . . . . 8 Figure 1-4. The activated methyl cycle and the derivation of DPD. . . . . . . . . . . . . . 11 Figure 1-5. The Lsr system found in most E. coli. . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 Figure 1-6. Lsr system effects biofilm development in multiple species. . . . . . . . . . .18 Figure 1-7. Possible bimodality arising from Lsr activity. . . . . . . . . . . . . . . . . . . . . . 24 Figure 2-1. Test Queries: lac Operon and Lsr System. . . . . . . . . . . . . . . . . . . . . . . . . 32 Figure 2-2. LMNAST heuristic. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35 Figure 2-3. Lac operon LMNAST hits overlaid onto phylogenetic distributions of different scopes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41 Figure 2-4. Coincidence heat map for lac operon LMNAST stringent search hits. . . 43 Figure 2-5. 2D similarity plot of lac operon LMNAST stringent search hits overlaid with attributed annotation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46 Figure 2-6. Lsr system LMNAST hits overlaid onto phylogenetic distributions of different scopes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49 Figure 2-7. Annotated 2D similarity plot for Lsr system LMNAST weak search hits. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51 Figure 2-8. Coincidence matrix for E. coli Lsr system LMNAST stringent search hits. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54 Figure 2-9. Coincidence matrix for E. coli Lsr system LMNAST extended window stringent search hits. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55 ix Figure 2-10. Phylogenetic distribution of Lsr at different phylogenetic scales using reconciled LMNAST results. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58 Figure 2-S1. Trackback plots for Lsr system LMNAST extended window stringent search hits. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64 Figure 2-S2. GC content demonstrates consistent spiking dip at intergenic region. . 68 Figure 3-1. LuxIR and Lsr QS activity intracellularly and intercellularly. . . . . . . . . 76 Figure 3-2. Lsr autoinduction of pure cultures leads to bimodal phenotype. . . . . . . . . 83 Figure 3-3. QS dynamics coupled with gliding during colony growth. . . . . . . . . . . . . 86 Figure 3-4. Cluster-disperse pattern from combination of Lsr and chemotaxis. . . . . . 88 Figure 3-5. Mixed culture simulations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89 Figure 3-S1 Comparison of solution for population with a single basal value versus a population with a unimodal distributed value of basal . . . . . . . . . . . . . . . . . . . . . . . . . 112 Figure 3-S2. Numerical solution for selected Lsr components and AI2-P. . . . . . . . . . 113 Figure 3-S3. Parameter sensitivity to the parameter, . . . . . . . . . . . . . . . . . . . . . . 114 Figure 3-S4. Dual ODE system simulation where second population has varied parameter values . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116 Figure 3-S5. Fraction of cell population QS activated decreases as the variation of the parameters Ksynth and VydgG increases. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118 Figure 3-S6. Comparison of results from single versus multiple finite difference elements to define environment. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120 Figure 3-S7. Congruence of solution from finite difference based modeling versus implicit solution of pure ODEs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122 Figure 3-S8. Measures of the difference between LuxIR and Lsr activation in the context of colony growth. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123 Figure 3-S9. Clustering of cells with lsr activity and AI-2 chemoattraction as measured by cell-cell distance. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126 x Figure 3-S10. Measures of the difference between different modes of motility when coupled with Lsr/AI-2 dynamics. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129 1 Chapter 1: Introduction 1 1.1 Background 2 Originally termed autoinduction 1,2, “quorum sensing” (QS) is an admitted 3 misnomer. The first description of a bacterial “quorum” clearly fell within the context of 4 cell concentration 3. Nonetheless, for the last two decades the popular equivalency of 5 “quorum” to a “sufficiently large population” has been rigorously reevaluated 3. With the 6 aid of new microfluidic schemes, researchers have shown that QS can be activated for 7 one or only a few isolated bacteria 4–6. With the benefit of careful consideration, others 8 have indicated that under some circumstances, QS might more appropriately be called 9 “diffusion sensing” or “efficiency sensing” 7–9. Yet others have argued for paring back 10 the usage of the term “quorum sensing” from an evolutionary point of view 10. 11 Indeed, QS studies have become home to a broad spectrum of bacterial 12 intraspecies and interspecies autoinduction or autoinduction-like signaling. This ready 13 adoption is probably attributable not only to the fact that QS is an apparently widespread 14 phenomenon, but could also be ascribed to the allure of the underlying paradigm: 15 unicellular organisms using self-secreted molecules to drive multicellular behavior. As 16 an additional point of interest, many QS-driven behaviors, such as toxin production and 17 biofilm formation, are tied to pathogenic phenotypes. The current count of molecules 18 generally considered autoinducers includes autoinducing peptide (AIP) from certain 19 Gram positive bacterial species, autoinducer-1 (AI-1 or acyl homoserine lactones 20 (AHLs)) from myriad beta- and alpha-proteobacteria, autoinducer-2 (AI-2 or (S)-4,5-21 dihydroxy-2,3-pentandione (DPD) and its interconvertible stereoisomers) from a great 22 2 array of bacteria, autoinducer-3 (AI-3, of unidentified composition) from E. coli, Cholera 23 autoinducer-1 (CA-1) and its analogs found in other Vibrio species, and Pseudomonas 24 Quinolone Signal (PQS). Notably, S-adenosyl methionine (SAM) lies directly upstream 25 of both AHLs and AI-2 synthesis, as noted in Figure 1-1. Serving as a methyl donor in 26 the activated methyl cycle, marked cellular abundance, and high reactivity may make it a 27 favored target for such repackaging. 28 Each QS signal is matched to one or more effector pathways that fall under one of 29 two broad categories, as depicted in Figure 1-2: two component regulatory systems 30 (TCRS) or direct transcription factor mediation. TCRS are typically composed of a 31 transmembrane receptor with histidine kinase functionality that activates a cytosolic 32 regulatory partner 11. For these systems, autoinducers remain extracellular when 33 generating a response. A separate category of QS activity is affected through the direct 34 mediation of transcription factor activity by autoinducer or modified autoinducer. 35 Occurring in the cytosol, this requires autoinducer internalization through either passive 36 or active means. Such cases include the activity of LuxIR found in numerous beta- and 37 alpha-proteobacteria 12 and Lsr systems found dispersed mainly among gamma-38 proteobacteria 13,14, respectively. 39 40 41 42 3 43 Figure 1-1. Synthesis of AHL and DPD (from 15). AHL and DPD share a common 44 upstream reactant, SAM, from the activated methyl cycle. 45 46 47 48 49 50 51 52 53 54 55 56 4 57 Figure 1-2. QS molecule activation of QS signaling pathway. A. Signaling molecules 58 activate the cell through receptor mediation at the cell surface, usually through a two 59 component regulatory system. B. Signaling molecules traverse the membrane and 60 interact with transcription factors directly or after slight modification. 61 62 63 64 65 66 67 68 69 5 Perhaps reflecting the complexity of their environment, many bacteria appear to 70 operate under the influence of multiple QS signal-effector units. Sometimes this appears 71 as topological redundancy or at least functional complementarity. In other cases, QS 72 signals appear to compete for influence, although this too could be a type of 73 complementarity, when discordant phenotypes are separately timed.16 74 1.1.1 QS in Vibrio harveyi 75 For instance, Vibrio harveyi QS is known to rely on at least three autoinducers: 76 3OHC4-HSL (an AHL), DPD, and a V. harveyi version of CA-1. In this instance, all 77 three autoinducers interact with separate cognate TCRS pairs, the signals of which are all 78 funneled into and filtered through the same overlapping negative feedback loops, 79 preventing premature activation under low autoinducer conditions and limiting 80 expression at high autoinducer concentrations, ultimately phosphorylating and activating 81 a single regulatory protein, LuxO.17,18 Despite this canalization, evidence suggests that 82 different autoinducers accumulate at different rates and help propel separate genetic 83 programs at distinct phases of growth 16, possibly triggering activity in combination with 84 distinct sets of transcription factors. 85 Although signal crosstalk between species can occur when AHLs are structurally 86 similar19 and is likely where the same molecule has been synthesized by different species 87 20, whether this necessarily represents QS qua QS is open to interpretation 10. As a 88 simplifying generalization, the activity of AHLs are commonly described as limited to 89 the species from which they are synthesized 21,22. For example, as far as is known, V. 90 harveyi’s AHL, 3OHC4-HSL 23, conforms to this paradigm. Furthermore, orthologs for 91 6 V. harveyi’s 3OHC4-HSL synthase and receptor, luxM and luxN respectively, are 92 phylogenetically confined to a small number of Vibrios 24–26, and produce distinct AHLs. 93 The limited scope of AHL signaling contrasts sharply with the more universally 94 produced AI-2—which is to say that DPD synthase (LuxS) homologs are found in myriad 95 species 27. While DPD exists as a specific chemical species, it spontaneously 96 interconverts between several distinct products, each possibly producing a different level 97 of QS activity 28. In V. harveyi, AI-2 interacts with the TCRS, LuxPQ 29, which 98 according to iterative protein BLAST searches is highly abundant among Vibrio and 99 Marinomonas species. 100 A particular derivative formed in marine environments where borate is abundant 101 is borated DPD 29. DPD is produced by practically all Vibrio species, and borated DPD 102 appears to be the cognate AI-2 molecule for those same species 30. 103 A third V. harveyi QS system involves cqsA and cqsS, a receptor-synthase pair 104 that is conserved widely but exclusively among Vibrio species. As with AHLs, the 105 resulting autoinducers (CA-1-like) are not known to have wide signaling efficacies 106 outside of the species by which they are produced 31. 107 Mediation of QS processes at the membrane is a distinctive characteristic of these 108 QS systems in comparison to other known QS machinery, and makes V. harveyi and the 109 Vibrio and Marinomonas species (all marine bacteria) that share these architectures 110 outliers. 111 Along the separate continuum of QS processes that feedback onto autoinducer 112 concentration, V. harveyi represents an intermediate in the QS landscape. On one end of 113 7 this spectrum exists QS systems like the canonical luxIR pair, where AHL synthase 114 expression is driven by autoinducer-activated transcription, producing a strong 115 intercellular positive feedback loop. On the other end is QS activation that is not only 116 dissociated from autoinducer synthesis but also drives uptake, producing a strong 117 negative intercellular feedback loop. 118 1.1.2 LuxIR QS 119 Widely utilized in synthetic biology, the LuxIR system is composed of an AHL 120 synthase, LuxI, and an AHL sensitive transcription factor, LuxR. Its operations are 121 depicted in Figure 1-3. Freely diffusing across the membrane, AHL extracellular 122 concentration is reflected intracellularly and vice versa. Given sufficient AHL with 123 which to complex, LuxR becomes less prone to degradation, accumulates, and promotes 124 the expression of both itself and LuxI, increasing AHL synthesis, thereby completing an 125 intracellular positive feedback loop 21. Furthermore, this sharp increase in AHL synthesis 126 leads to a corresponding increase in local extracellular concentration, coordinating the QS 127 activation of neighboring cells which in turn influence their own neighbors, generating 128 positive feedback at a multicellular scale. This system core and variations thereof 129 represent the dominant pathway by which AHLs influence downstream phenotypes. 130 8 131 Figure 1-3. LuxIR activity at single cell and multicellular hierarchies. A. AHL freely 132 traverses the membrane and stabilizes the transcription factor, LuxR. This activates a 133 positive feedback loop, increasing LuxR expression as well as that of LuxI, the AHL 134 synthase. B. This positive feedback operates at the multicellular level as well. 135 136 137 138 139 140 141 142 143 144 9 LuxIR homologs are widespread throughout proteobacteria. As of 2008, 26% of 145 sequenced bacterial genomes contained a complete set of homologs 12. Among them are 146 the pathogens Yersinia pestis, Agrobacteria tumefaciens, and Pseudomonas aeruginosa. 147 Many species contain multiple and distinct luxIR pairs 12, which were likely acquired 148 independent of each other even if their signaling converges downstream 32. 149 Effective at organizing population expression, luxIR systems have often been 150 repurposed toward applied ends. Initial efforts usually involved adding exogenous genes 151 under the luxR promoter, such as an enzymatic degrader of AHL to create an oscillating 152 signal 33 or chloramphenicol to limit population density despite sufficient nutrient in the 153 environment 34. In certain cases, luxIR QS was used to ensure complete induction within 154 a population 35. More advanced designs included separating luxIR sensing and 155 autoinducer production capacities into multiple bacterial carriers whether for synthetic 156 biology or therapeutic ends 36,37. 157 In addition to complete luxIR pairs, most species contain more luxR homologs 158 than luxI homologs.12 Possibly, at least some of these unmatched luxR homologs are not 159 redundant, but serve as a means of testing for different species’ AHLs. This is the case 160 for Rhizobium leguminosarum bv. viciae which depends upon exogenous N-(3-hydroxy-161 7-cistetradecenoyl)-L-homoserine lactone (3-OH-C14:1-HSL) to induce conjugation 38. 162 The luxR homolog, sdiA, found in E. coli, Salmonella, and related enterobacterial species 163 39, also appears to support such a role. While these species do not synthesize any AHL, 164 they nevertheless respond to several AHLs through SdiA 40–42. As might be expected of a 165 quorum sensing regulator, sdiA is at least in some species tied to biofilm activity 43, 166 although the extent of its reported signaling has varied depending on whether the study 167 10 involved a knockout 44, plasmid based overexpression 42, or reporter fusion into the 168 chromosome 45. 169 1.1.3 AI-2 QS 170 Perhaps even more broadly distributed than AHL based associated QS is AI-2 171 driven QS. However, this is not without caveats. For whereas many bacteria have AHL 172 receptors without producing the specific AHL, many bacteria appear to produce AI-2 173 without an apparent ability to perceive the signal. This inversion has previously led some 174 to suggest that AI-2 may represent a metabolic by-product rather than a true signaling 175 molecule. Indeed, LuxS in combination with Pfs are an integral part of regenerating 176 homocysteine after methyl donation as part of the activated methyl cycle, as seen in 177 Figure 1-4 46. This pathway serves as an alternative to S-adenosyl homocysteine (SAH) 178 hydrolase (SahH). While eukaryotes use the SahH pathway exclusively, the proportion 179 of bacteria expressing sahH is roughly split with that utilizing the luxs/pfs pathway 46. 180 With only the rarest exception are luxS and sahH ever coincident to the same genome 181 27,47. Importantly, bacteria containing sahH are unable to produce AI-2 except by 182 conversion from ribulose-5-phosphate 48,49. Whether the flux from this secondary 183 synthesis pathway is sufficient to effect even limited QS operations remains unknown 49. 184 Still, while AI-2 production may not be universal, luxS is nonetheless well represented 185 among bacterial genomes, and like AHL, AI-2 signaling is not restricted to the bacteria 186 from which it is synthesized. Moreover, whether various combinations of exogenous in 187 vitro synthesized 27 or purely synthetic 50 AI-2 and exogenous sahH expression can 188 rescue luxS mutant phenotypes provides an interesting point of study. 189 11 190 Figure 1-4. The activated methyl cycle and the derivation of DPD (from 51). AI-2 is an 191 umbrella term for a number of spontaneously cyclizing DPD derivatives. DPD itself is a 192 byproduct of the activated methyl cycle during the regeneration of homocysteine, at least 193 in some bacteria. Most other cells with an activated methyl cycle utilize the alternative, 194 SahH, which bypasses the production of DPD. 195 196 197 12 Less phylogenetically dispersed than AI-2 synthesis, receptors for the molecule 198 are nonetheless found in numerous bacteria. Predominantly counted among bacteria with 199 receptors are gamma-proteobacteria, for which the luxS transgenomic signal appears to be 200 the most monophyletic 46,52. This is partly attributable to LuxP receptors found among 201 numerous Vibrio and Marinomonas species, as previously discussed. Among non-Vibrio 202 and non-Marinomonas species, the only other known AI-2 QS receptor is the periplasmic 203 receptor protein, LsrB, the binding component of an ABC type transporter. Originally 204 identified in Salmonella along with the rest of the LuxS regulated (Lsr) system 53, LsrB 205 has been shown through reciprocal BLAST searching to be most prevalent, but not 206 universal in, nor restricted to Pasteurellaceaes and Enterobacteriales 13. Just as Vibrios 207 and Marinomonas species exist in a shared environment, it could be suggested that LsrB 208 has been transferred horizontally among bacteria sharing the same environment 14. 209 In its sum, the Lsr system acts as a sugar-like importer system, composed of 210 divergently arranged operons that share an intergenic region, lsrRK and lsrACDBFGE in 211 Salmonella 53,54. lsrE is absent in E. coli, which otherwise has a nearly exact duplicate 55–212 57. LsrR belongs to the deoR family of transcriptional repressors, interacting with the 213 intergenic region through a helix-turn-helix motif, keeping system expression to a 214 minimum when sufficient substrate is unavailable. LsrA is the nucleotide binding 215 component of the ABC type transporter, binding and using ATP to drive AI-2 import. 216 LsrC and LsrD are transmembrane proteins comprising a pore for AI-2 between the 217 cytosol and the periplasm. As described previously, homology indicates that LsrB binds 218 AI-2 in the periplasm, where the complex joins to the transmembrane components. Once 219 inside the cytoplasm, AI-2 is phosphorylated by LsrK. Notably, lsrK appears to have its 220 13 own constitutive promoter, albeit one that does not drive significant expression 56. 221 Phosphorylated AI-2 (AI2-P) de-represses system expression by destabilizing LsrR’s 222 interaction with DNA. AI2-P is also broken down into multiple products by LsrF and 223 LsrG. The exact function of lsrE is unknown, although it is commonly alternatively 224 annotated as ribulose-3-phosphate epimerase. This, and its operon position suggests it 225 may operate on AI2-P 51. Unlike lsrF and lsrG, however, it fails to effect apparent AI2-P 226 levels 54. 227 The Lsr system is also affected by multiple factors outside its apparent regulon. 228 For example, cAMP is required for system expression 55. Additionally, AI-2 export 229 requires the transmembrane protein YdgG, without which AI-2 remains sequestered 230 within the cell 58. YdgG appears to belong to a well conserved superfamily of exporters 231 that have no other ascribed function 59. Finally, it is believed that an additional low flux 232 importer pathway is required for import when the Lsr system is inactive 55,60. 233 A few additional proteins have also been shown to mediate AI-2 activity at the 234 cell membrane. These include RbsB from the ribose importer system in Aggregatibacter 235 actinomycetecomitans 61, TlpB which allows Helicobacter pylori to be chemotactically 236 repelled by AI-2 62, and the PTS system in E. coli 63. While TlpB is a broad array 237 chemoreceptor, both RbsB and PTS are involved in the import of sugars across the 238 membrane. RbsB was shown to compete with LsrB for AI-2 and that it was required for 239 Lsr activation 61. In a certain sense, however this is unsurprising insofar as many high 240 affinity transporters for one monosaccharide act as low affinity transporters for related 241 monosaccharides. For example, GalP imports glucose and galactose at high rates, and 242 14 243 Figure 1-5. The Lsr system found in many E. coli. The ABC-type transporter and LsrK 244 serve as positive feedback elements, working serially to produce AI-2P. LsrR and 245 LsrF/G, on the other hand, serve as negative feedback elements, encouraging system 246 repression or discouraging system de-repression respectively. Outside of these elements 247 exist the transporters YdgG (an exporter) and a low affinity alternative importer pathway 248 (consisting, at least, of the PTS system). 249 250 251 252 253 254 15 lactose and fructose at lower rates when both of its primary targets are absent 64,65. This 255 also obtains for ALBP, which primarily transports allose across the membrane but also 256 doubles as a low flux pathway for ribose 66. Both allose and ribose contain aldehyde 257 functional groups. That both ribose and AI-2 are both 5 carbon molecules lends credence 258 to the possibility of such a secondary role for RbsB. While RbsB is at least indirectly 259 involved in Aggregatibacter actinomycetemcomitans’s AI-2 response 67, the regulatory 260 pathway is unknown. Possibly, the rbs system comprises the alternative low flux 261 importer affecting Lsr response, but this remains a subject of further research. 262 Interestingly, a system annotated as rbs in Haemophilus influenzae also appears to 263 regulate AI-2 influx. Additionally, this rbs’s expression is also controlled by AI-2. In 264 the reported studies, none of the pentose sugars tested, including ribose, induced 265 competitive inhibition, however, indicating that this may simply be a case of improper 266 annotation 68, serving as a reminder of possible annotation biases associated with 267 precedence. Many of the same ideas regarding AI-2 cross-reactivity with rbs apply to the 268 PTS system which serves as a flux pathway for a variety of sugars, and upon which Lsr 269 signaling is dependent in E. coli 63. 270 1.1.4 Lsr regulon 271 The exact pathway by which signal is transduced by the Lsr system appears to 272 vary from species to species. Two separate studies have assessed the nature and number 273 of potential LsrR binding sites. According to ChIP-CHIP analysis, the LsrR regulon is 274 limited to the Lsr system itself in Salmonella Typhimurium 69. In E. coli, the LsrR 275 regulon is ostensibly more expansive. While the footprinting differential between lsrR 276 knockouts and wildtype cells is strongest near the lsr intergenic region, additional 277 16 binding sites include regions near yegE, mppA, and yihF 70. Both yegE and mppA assist 278 in the transition from exponential growth toward a sessile lifestyle. A probable 279 diguanylate cyclase, yegE appears to help coordinate the concomitant downregulation of 280 flagellar genes and upregulation of curli 71. Separately, mppA is involved in recycling 281 membrane associate peptides 72. 282 Additional means by which Lsr may influence downstream phenotypes exist in 283 the form of LsrF’s and LsrG’s enzymatic outputs. While LsrF is believed to be an 284 aldolase, its specific products remain unknown 73. LsrG is thought to act as an isomerase 285 under anaerobic conditions to produce 3,3,4-trihydroxy-2-pentanone-5-phosphate 74, 286 whereas it cleaves AI2-P under aerobic conditions to produce 2-phosphoglycolic acid 287 (PG) and an unidentified 3 carbon molecule 75. PG acted upon by the phosphatase gph 288 reenters metabolism as glycolic acid 76. Whether any of these products affects 289 downstream phenotypes remains unknown. 290 1.1.5 Known Lsr Associated Phenotypes 291 The reported effects of LsrR mutation and other Lsr system modifications mirrors 292 the ChIP-CHIP results. LsrR did not have an apparent impact on the measured 293 phenotypes, invF and flagella gene expression, in Salmonella Typhimurium except when 294 luxS was also deleted or when LsrR was overexpressed 77. This suggests that LsrR 295 binding to DNA may be easily destabilized by the presence of AI2-P in Salmonella, 296 perhaps pointing to a reason why LsrR’s direct regulon may be limited in this species. 297 On the other hand, in E. coli, deletion of either lsrR or lsrK led to marked inhibition of 298 biofilm development 60 (Figure 1-6a). Other experiments involving luxS knockouts with 299 17 homocysteine and SAM controls indicated that exogenous DPD restored biofilm height 300 while not fully complementing biomass accumulation 78. Interestingly, biofilm 301 phenotype was also inhibited by lsrR deletion in Aggregatibacter 302 actinomycetemcomitans, a bacteria common to the oral cavity 79. In this case, although 303 lsrRK knockout led to more dramatic changes than lsrR deletion alone, statistically 304 significant changes arising from lsrK deletion remained unidentified, leading to a 305 suspicion that an additional kinase may be working on cytosolic AI-2 79 (Figure 1-6b). 306 This small cohort of studies indicates that the Lsr system can influence population scale 307 phenotypes in a manner expected of QS systems. 308 1.1.6 Broader AI-2 Associated Phenotypes 309 In general, QS drives behaviors that work most efficiently when orchestrated at 310 multicellular scales 80. Along with swarming and sporulation, among the more strikingly 311 cooperative of QS guided behaviors is biofilm formation. In addition to the apparent 312 cooperativity involved, biofilms are also known for their protective role, which lends to 313 bacteria’s robustness against environmental insults (e.g. oxidation and antibiotics) and 314 also plays a role in refractory infections. This is a common theme in QS signaling, which 315 appears to influence numerous aspects of virulence including motility switching, bacterial 316 adhesion, invasion, and toxin production. More broadly, AI-2 signaling might be 317 expected to control numerous aspects of bacterial interactions in natural consortia. 318 In addition to the species previously mentioned, AI-2 in particular has been 319 shown to influence biofilm formation and other cooperative behaviors in many bacteria, 320 often in a manner similar to that in E. coli, where, as discussed previously, exogenous 321 18 322 Figure 1-6. Lsr system effects biofilm development in multiple species. A. As adapted 323 from 60, E. coli biofilm development appears stalled upon deletion of either lsrR or lsrK. 324 B. The same appears to be true for A. aggregatibacter, except the effect from lsrK 325 mutation was not statistically significant. Dual deletion resulted in a sparser biofilm, and 326 complementation on a plasmid restored that defect. 327 328 329 330 331 19 DPD partially complemented luxS deletion. This inexact restoration may be attributable 332 either to the dual signaling and metabolic role played by LuxS or to the transport 333 limitations of exogenous DPD supplementation at both macroscales and microscales (i.e. 334 around biofilm architecture or across the cell membrane). 335 A couple studies have focused on bacterial species synthesizing AI-2 and 336 containing Lsr homologs, without explicitly studying Lsr based expression. DPD 337 complemented luxS mutation in Photorhabdus luminescens for some differentially 338 expressed genes to an unspecified degree while others were unaffected. Fully or partially 339 restored expression was found in virulence genes such as hemolysin and pyocin, 340 oxidative response genes such as msrB (Peptide methionine sulfoxide reductase B) and 341 uvrC (Exonuclease ABC subunit C), and motility related genes such as flhC (Flagellum 342 biosynthesis transcription activator). In this report researchers were additionally unable 343 to duplicate DPD induced changes with homocysteine supplementation.81 In Bacillus 344 cereus, exogenously added AI-2 affected biofilm growth in a closely titrated manner, 345 initially increasing biofilm accumulation but also inhibiting biofilm growth when 346 exposed to a higher concentration within the same order of magnitude 82. 347 A separate group of bacteria producing AI-2 but lacking a known AI-2 receptor 348 also appear to have AI-2 sensitive phenotypes. Exogenously added DPD increased 349 adhesion in a dose dependent fashion in both luxS mutant and wildtype Actinobacillus 350 pleuropneumoniae populations 83. In Streptococcus pneumonia, researchers were able to 351 demonstrate partial reestablishment of biofilm biomass to wildtype levels in a luxS 352 mutant treated with exogenous DPD 84. In Streptococcus epidermidis, exogenous AI-2 353 complementation restored PSM (pro-inflammatory phenol soluble modulin peptides) 354 20 expression in addition to that of many metabolic genes that were differentially expressed 355 in a luxS knockout compared to a WT strain 85. In Streptococcus intermedius, AI2 356 exogenously added to luxS mutants again produced a distinct biofilm phenotype 357 compared to WT and luxS mutants in both the presence and absence of antibiotics. 358 Moreover, AI-2 conferred some ability to form a basal level of biofilm in the face of 359 several classes of antibiotic.86 Among these reports, one involving Borrelia 360 burgdorferi serves as a particularly illuminating example. For this species and in many 361 of its relatives, the activated methyl cycle is incomplete. In particular, B. burgdorferi 362 cannot generate methionine from homocysteine. Nonetheless, AI-2 is produced through 363 the LuxS/Pfs pathway, and moreover, DPD can complement luxS deletion, restoring 364 expression of vlsE and erp, both of which translate into proteins that mediate the host-365 pathogen interaction.87 In addition to its ability to chemorepel from AI-2, H. pylori also 366 appears to regulate multiple flagellar genes with AI-2, as AI-2 complements luxS deletion 367 and also enhances these flagellar gene expression levels when added to wildtype cells. 88 368 A number of other bacteria that do not synthesize AI-2 also respond to 369 exogenously supplied DPD. Ranking among these is Sinorhizobium meliloti, which 370 contains an Lsr homolog 14 and can deplete AI-2 from the extracellular environment 371 without apparent consequence to colonization or growth phenotypes 89. Pseudomonas 372 aeruginosa expression of virulence genes is influenced by exogenous AI-2 90. Finally, 373 DPD was shown to influence biofilm production in a dose responsive fashion via the 374 induction of an oxidative stress response pathway in Mycobacterium avium 91. Unlike S. 375 meliloti neither P. aeruginosa nor M. avium has an identified AI-2 response pathway. 376 377 21 1.1.7 QS in an Ecological Context 378 From an evolutionary point of view, QS and AI-2 based-QS in particular 379 represent a peculiarity due to perceived, or at least potential, cooperation between 380 nonrelated bacterial species. By way of illustration, pure culture QS signaling that 381 encourages biofilm formation is an entirely suitable method by which to organize 382 population-wide expression of public goods in an isolated setting. In such instances, QS 383 based cooperation can be attributed to kinship.92 When considering QS related mutations 384 and the likely mixture of genetic backgrounds and species within a bacterial consortium, 385 however, the exact manner in which evolutionary pressures inform and maintain QS 386 signaling becomes less straightforward. In even moderately more complex 387 circumstances, questions of which bacteria produce signal, which bacteria respond to 388 signal, and which bacteria benefit from any resulting product can quickly overwhelm the 389 boundaries of current understanding. 390 391 392 393 394 395 396 397 22 1.2 Research Motivation 398 Lsr based QS in E. coli represents a practical model for studies herein. Among 399 the better characterized of the Lsr systems 49,55–58,60,63 and effecting virulence phenotypes 400 60,78, the study of such a system is of natural interest for a diverse generalist bacteria like 401 E. coli, certain strains of which are a common source of food supply contamination. 402 Arising from such concerns, the identification of strains containing Lsr system homologs 403 and how Lsr based QS might play a role in infection processes among these strains is of 404 interest. Yet even more broadly, it is important to identify other species in which the Lsr 405 system may play a signaling role. While much research has been conducted to identify 406 AI-2 responsive bacteria, sometimes through the Lsr system, questions of 407 generalizability, while somewhat epistemological in nature, are also useful when answers 408 can be had. By the identification of useful homology, and determining how homologs are 409 constituted, one may begin a line of inquiry into questions of generalizability. 410 The phylogenetic dispersion of AI-2 receptors has been assessed as a means of 411 determining which bacteria might be responsive to AI-2 as an external signaling stimulus 412 13. The effectiveness of such searching is largely dependent upon the increasingly rapid 413 pace of whole genome sequencing, and while many such receptors have been identified 414 there remain numerous AI-2 responding bacteria with unknown receptors. While it is 415 possible additional sugar transporters moonlight as high affinity AI-2 transporters, sugar 416 transporters acting as a lone pathway for AI-2 import has yet to be demonstrated in any 417 species. In general, the perception of AI-2 is of natural interest, as some cells may 418 recognize it at lower concentrations than others, etc; and here, perception is meant to not 419 only indicate reception, but also encompasses subsequent gene expression changes. Here 420 23 we wish to only consider possible Lsr gene expression changes, while remaining aware 421 that the Lsr system may effect a broader regulon. Toward this end the identification of 422 extant Lsr genes, operon organization 93, and the phylogenetic distribution of variants 423 among Lsr homologs is of interest. 424 A second cohort of studies described herein were initially motivated by 425 experiments indicating that stable bimodal Lsr system expression may arise in pure 426 cultures 94 (Figure 1-7a). While it has been suggested that a transient bimodality may 427 arise as a consequence of intracellular signaling topology 95, a more permanent 428 population bifurcation may also develop from hyperlocal competition for AI-2 between 429 cells, as depicted in Figure 1-7b. Bimodal expression within a pure culture would 430 represent a built-in population of QS cheats, and depending on the expected benefit could 431 speak to issues about free riders and other evolutionary conundrums. 432 The picture is further complicated for bacteria that are chemoattracted to AI-2, 433 such as E. coli 96,97. Although QS generally informs population scale behaviors, 434 activation of Lsr and the resulting net recompartmentalization occur on time and spatial 435 scales that intersect with chemotaxis. As a means of isolating this confluence, a mixed 436 finite difference-agent based approach can be used to model emergent behaviors 98. 437 Using such a platform, contextualized QS expression between two competing populations 438 with different genotypically mixed signal/reception competency can be interrogated. 439 Specifically, as a possible consequence of its universality, the literature suggests that AI-440 2 signaling within a consortia is potentially quite complex. How QS operates in an 441 environment with multiple systems operating concurrently remains an open question. 442 Thus, in addition to investigating QS by pure cultures, this framework can also be used to 443 24 444 Figure 1-7. Possible bimodality arising from Lsr activity. A. FACS results adapted from 445 94 suggest that a bimodality develops over time. B. A diagram of a possible mechanism 446 whereby bimodality could develop, as extracellular negative feedback operates 447 hyperlocally. 448 449 450 451 452 453 454 455 456 457 458 25 investigate basic QS systems acting concurrently in both time and space. Specifically, 459 how different Lsr system variants might compete with one another can be investigated. 460 Essentially, in addition to determining which species and strains might be 461 responsive to AI-2, there exists additional opportunity to ask how they might be 462 responsive to AI-2. Research presented herein takes a closer look at the phylogenetic 463 distribution of Lsr system homologs and the composing members of its homologs. We 464 then consider more closely how the Lsr system possibly produces bimodal expression 465 through the mathematical modeling of two competing populations. We further consider 466 what the ramifications of system behavior might be in the context of motility, and how 467 multiple bacteria with different Lsr system variants might compete for AI-2 with one 468 another. 469 470 471 472 473 474 475 476 477 478 26 1.3 Global Objective, Global Hypothesis, and Specific Aims 479 The global objective of this dissertation is to further investigate the nature of the Lsr 480 system, using E. coli K-12 W3110’s Lsr system as a starting point, both interrogating 481 how W3110’s Lsr homologs compare to each other and how they compare to other QS 482 systems. 483 Global Hypothesis: E. coli’s Lsr system is one among a spectrum of Lsr systems, the 484 diversity of which bears signaling consequences both in pure cultures and mixed 485 consortia. 486 Specific Aim 1: Find E. coli Lsr system homologs. 487 Specific Aim 2: Determine if the known components of the Lsr system, along with 488 auxillary proteins can lead to stable, bimodal Lsr activation. 489 Specific Aim 3: Assess how Lsr’s recompartmentalization dynamics may differentiate it 490 from other QS systems within the context of population activation, emergent 491 phenomenon, and competition between bacteria containing different Lsr systems. 492 493 494 495 496 497 27 1.4 Dissertation Outline 498 Chapter 2 describes a novel algorithm to find homologs of modular networks and an 499 analysis of the results when using the E. coli K-12 Lsr system. This is contrasted to the 500 results for the lac system from the same strain. 501 Chapter 3 describes the development of a set of ODE’s that describes the development of 502 bimodal system expression given parameter variation. 503 Chapter 4 describes the development and results of a finite difference-agent based model 504 contrasting LuxIR activation to Lsr activation within the context of emergent behaviors 505 arising from the confluence of QS and varying motility modes. 506 Chapter 5 summarizes the previous chapter’s work and indicates challenges to future 507 work, denoting the larger implications and significance of the work. 508 509 510 511 512 513 514 28 Chapter 2: Comparison of Homolog Identification for the sugar 515 importing Lac System and the QS Lsr System from E. coli K-12 W3110 516 517 2.1 Abstract 518 Bacterial cell-cell communication is mediated by small signaling molecules known as 519 autoinducers. Importantly, autoinducer-2 (AI-2) is synthesized via enzyme LuxS in over 520 80 species, some of which mediate their pathogenicity by recognizing and transducing 521 this signal in a cell density dependent manner. AI-2 mediated phenotypes are not well 522 understood however, as the means for signal transduction appears varied among species, 523 while the AI-2 synthesis process appears conserved. Approaches to reveal the recognition 524 pathways of AI-2 will shed light on pathogenicity as we believe recognition of the signal 525 is likely as important, if not more, than the signal synthesis. LMNAST (Local Modular 526 Network Alignment Similarity Tool) uses a local similarity search heuristic to study gene 527 order, generating homology hits for the genomic arrangement of a query gene sequence. 528 We develop and apply this tool for the E. coli lac and LuxS regulated (lsr) systems. Lsr 529 is of great interest as it mediates AI-2 uptake and processing. Both test searches 530 generated results that were subsequently analyzed through a number of different lenses, 531 each with its own level of granularity, from a binary phylogenetic representation down to 532 trackback plots that preserve genomic organizational information. Through a survey of 533 these results, we demonstrate the identification of orthologs, paralogs, hitchhiking genes, 534 gene loss, gene rearrangement within an operon context, and also horizontal gene transfer 535 (HGT). We also found a variety of operon structures that are consistent with our 536 29 hypothesis that the signal can be perceived and transduced by homologous protein 537 complexes, while their regulation may be key to defining subsequent phenotypic 538 behavior. 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 30 2.2 Introduction 555 Comparing prokaryotic whole genome sequences to identify operons is a mature 556 area of research 99–102. Orthologous operon identification can imply a secondary degree 557 of relation between components, reaffirming Clusters of Orthologous Groups (COG) and 558 other assignments of function as well as suggesting essentiality 103. This conservation of 559 components also speaks to the conservation of signaling capacity in orthologous modular 560 signaling operon-based units. That is, we are interested in ascertaining the genetic 561 modularity of signal transduction processing, in particular those that operate within 562 known, putative regulons. Drawing partly on previous work investigating microsynteny 563 and gene neighborhoods 101,104,105, we developed a general similarity search approach, we 564 call a Local Modular Network Alignment Similarity Tool (LMNAST). LMNAST applies 565 a BLAST-like heuristic to gene order and arrangement. Resultant search hits help 566 capture the conservation and phylogenetic dispersion of a given query modular network. 567 Using, as queries, contiguously abutting genes of prokaryotic modular signaling 568 networks, LMNAST identifies and scores hits based on the minimum number of frank 569 mutations in gene organization needed to arrive at a given putative system homolog 570 starting from the query. Here, homology refers to similarity in relative gene order and 571 relative transcriptional direction, after nucleotide level threshold filtering of gene 572 elements based on BLAST 106 E-value. 573 For the purpose of evaluation, two small modular systems were used as test 574 inputs: one was the E. coli lac system and the other was the LuxS regulated (Lsr) system. 575 In some ways, the two systems are quite similar (Figure 2-1). Both import and 576 31 catabolize the small molecules that induce system expression. For the lac system, this 577 small molecule is, of course, lactose. For the Lsr system, the small molecule is 578 autoinducer-2 (AI-2). AI-2 is a signaling molecule common among at least eighty 579 bacterial species 27. As mediated either through the Lsr system or LuxPQ, bacteria are 580 believed to use AI-2 to guide population based phenotypes, a phenomenon termed 581 quorum sensing 54. LuxPQ is a histidine kinase two component system, the regulon of 582 which is distinct from Lsr and is not considered further. Lsr is an interesting query 583 because its distribution should help elucidate its putative, modular quorum sensing 584 function 27 and because the known homologs differ in gene organization.54,55,82 585 586 587 588 589 590 591 32 592 Figure 2-1. Test queries: lac operon and Lsr system. A The lac operon is composed of 593 beta-galactosidase (LacZ), the lactose importer (LacY), and a beta-galactoside 594 transcetylase (LacA). Upstream of the operon, the operon repressor (LacI) is expressed 595 in a co-directional orientation. The primary function of the lac system is as a regulated 596 importer/processing unit. Lactose brought in through LacY is converted into allolactose 597 or hydrolyzed into glucose and b-galactose. Both reactions are catalyzed by LacZ. 598 Allolactose then acts to release the repression of the system by LacI. B The Lsr system is 599 composed of two divergent operons. One operon consists of an AI-2 kinase, and a system 600 repressor. The other operon consists of an AI-2 transporter and phospho-AI2 (AI2-P) 601 processing genes. Contextual system behavior is partly governed by separately regulated 602 parts including an alternative importer 61, an exporter 58, and the AI-2 synthase gene. 603 Relative to the canonical lac system, the Lsr system is complicated by the fact that the 604 cell synthesizes, exports, and imports AI-2, and by the negative regulation associated 605 with the divergently arranged structure. AI-2 exported by a mechanism involving YdgG 606 traverses the outer membrane through a porin and enters the periplasmic space. Through 607 the ABC-type importer, LsrACDB, AI-2 is then transported back into the cytosol. Once 608 33 there, AI-2 is phosphorylated by LsrK. This phosphorylated form (AI2-P) de-represses 609 the lsr system and is catabolized by LsrF and LsrG into separate downstream products. 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 34 2.3 Methods and Algorithm 627 As previously indicated, LMNAST evaluates nucleotide records for similarity to a 628 query network using a BLAST-like heuristic. Specifically, it captures the gene order of 629 query networks as a string of characters. Queries are therefore not necessarily restricted 630 to defined networks insofar as any gene ordering may be a query. A standard heuristic of 631 penalties for various rearrangements of orthologous system is employed. For searches 632 described herein, a loose threshold was used to generate an exhaustive set of hits. The 633 overall scheme is depicted in Figure 2-2. The program itself is available at 634 http://www.bentley.umd.edu. 635 636 35 637 Figure 2-2. LMNAST heuristic. LMNAST operates in a BLAST-like manner, using the 638 results of BLAST searches themselves as a curated database. 639 640 1. For each member of the query, in any nucleotide record, a homolog’s membership to a 641 character type is assigned by scoring below a specified BLAST E-value threshold. Genes 642 assigned to characters are highlighted blue. Genes without sufficient homology to any 643 character are represented by dashed boxes. 644 2. Sufficiently long stretches of adjacent characters are identified as seeds (red). 645 36 3. Sufficiently proximal characters are connected to seeds or seeds are connected to each 646 other when at a base pair distance < d. 647 4. Rearrangements, losses, and deletions are scored according to a standard similarity 648 heuristic. Noncontinuous elements are dropped iteratively until a maximum score is 649 achieved, arriving at… 650 5. An LMNAST hit. 651 652 653 654 655 656 657 658 659 660 661 662 663 37 2.3.1 Input 664 Input consisted of an ordered list of gene elements (for example, lacIZYA). For 665 each gene element a BLAST result file was generated using tblastn to search the nr/nt 666 database for hits with E-values less than 0.1, narrowing the search space. Each BLAST 667 hit was assigned a character corresponding to the gene element queried. BioPerl 107 was 668 used to query Genbank databases and process data from retrieved files. Nucleotide 669 records with sufficiently proximal characters were investigated further. 670 2.3.2 Scoring Heuristic 671 The degree of similarity between a putative hit and a query was evaluated 672 according to the number of deletions, insertions, and rearrangements required to generate 673 the putative hit using the query as a starting point. Intra-hit gene duplications were 674 disallowed as a simplification. Consequently, deletion could be noted by character type 675 inclusion. Insertions of uncharactered elements between gene homologs were scored 676 according to an affine gap rule whereby a portion of the deduction was scaled to the 677 insertion length. Rearrangements refer to altered relative order and relative gene 678 direction. Changed relative direction was only considered when relative order was 679 maintained. When this criterion was satisfied, relative order was evaluated in terms of 680 adjacent homolog distance, disregarding insertions and deletions. For each such 681 structural dissimilarity there was a standard deduction in score. Noncontiguous elements 682 were dropped iteratively until a maximum score was reached for each putative hit. When 683 putative hit versions elicited equal scores in the same round, the version of the hit with 684 38 the most characters was retained. Putative hits with scores greater than zero were 685 retained. 686 2.3.3 Weak and Stringent Criteria 687 For evaluation purposes and to find a suitable balance between false positives and 688 coverage completeness, each test query was run under both weak and stringent 689 conditions. Stringent criteria searches assumed accurate annotation. Contrarily, weak 690 criteria did not require genes to lie within the annotated coding sequence. Moreover, 691 characters annotated as “pseudo” or bounded outside “gene” annotation were accepted as 692 homologous characters. Weak criteria searches also allowed multiple genes to co-exist 693 within the same annotation. Additionally, as a concession to the possibility of longer 694 range interactions between genes, reduced gap penalties were used in weak criteria 695 searches. Results described herein were derived using a gap penalty of 1 and 2 with an 696 extension penalty of 0.3 and 1, for weak and stringent criteria searches respectively. 697 2.3.4 Ancillary LMNAST Search Tools 698 Mean element homology (meH) is a normalized, ancillary measure of string 699 similarity as evaluated by BLAST. Useful for contrasting BLAST results to LMNAST 700 hits, meH was calculated by normalizing each gene homolog’s bit score to the maximum 701 bit score for the entire corresponding BLAST result with a background subtraction of the 702 minimum bit score. These normalized bit scores were then averaged for all gene 703 elements within an LMNAST hit. A score of one indicates exact likeness whereas zero 704 indicates the least degree of similarity. 705 39 Also, widening the query beyond the system of interest to include a nominal 706 number of flanking genes, here termed “extended window searching,” afforded additional 707 contextualization of LMNAST hit results. 708 Finally, in evaluating certain low homology hits, nonscoring synonyms were 709 used. Nonscoring synonyms are elements with equivalent gene annotation but 710 insufficient homology according to the initial E-value filter. This is somewhat analogous 711 to replacement in blastp. 712 713 714 715 716 717 718 719 720 721 722 723 724 40 2.4 Results 725 2.4.1 E. coli K-12 W3110 lac Operon Query 726 We began evaluation of LMNAST by searching for the well characterized E. coli 727 lac operon. Specifically, the E. coli lac genes lacI (BAE76127), lacZ (BAE76126), lacY 728 (BAE76125), and lacA (BAE76124) (spanning bp 360473 to 366734 of the Genbank 729 nucleotide record AP009048) were used as a query. The stringent criteria search yielded 730 fewer hits than the corresponding weak criteria search (189 vs. 236). Of the hits derived 731 from the stringent criteria search, complete and perfectly arranged lac systems were 732 found in 26 unique E. coli strains and S. enterica arizonae serovar 62:z4,z23 (meH 0.8), 733 the only Salmonella enterica serovar represented among all lac system hits, in keeping 734 with its significant divergence from other serovars 74. A representation of E. coli hits in a 735 phylogenetic context is available in Figure 2-3a. The average meH (0 < meH < 1) for 736 these complete systems was 0.98. An extended window query with five additional genes 737 on either side of the original search frame, revealed eight complete systems with a 738 hitchhiking, proximal cytosine deaminase after losing all other proximal genes. Only one 739 system with all four characters was entirely removed from the original query’s proximal 740 gene set, suggestive of negligible stability for the canonical system outside of a highly 741 limited phylogenetic domain. 742 An additional 28 hits were bereft one lac system character (average meH 0.74). 743 In all but three of these cases that missing gene was lacA. Of the hits without lacA, ten 744 had an additional frank structural change to a divergent expression pattern originating 745 41 746 Figure 2-3. lac operon LMNAST hits overlaid onto phylogenetic distributions of 747 different scopes. The larger, bolded leaves represent species that contain lac operon 748 homologs, whereas the grayed italicized leaves were completely bereft. A An E. coli 749 specific phylogenetic tree as adapted from 102, wherein all genes from the core E. coli 750 were used to construct a consensus tree. Among the species/strains represented here, lac 751 system homologs were absent from certain Shigella. Additionally, BW2952, SE11, IAI1, 752 HS, E243227A, CFT073, K-12 DH10B, and 11 of 18 O:H serotyped strains contained 753 truncated systems. Uniquely, CFT073 retains lacA, while missing lacI in an otherwise 754 preserved lac system structure. B The phylogenetic dispersion of the lac system is 755 mostly limited to Escherichia and proximal species, as seen in the 16s based tree adapted 756 from 108. Bolded leaves indicate the presence of the lac system in at least one strain. 757 758 42 between lacI and lacZ characters (e.g. in E. cloacae), likely increasing system expression 759 sensitivity to lacI repression in these cases 93. Surprisingly, in other instances, extended 760 window searching revealed the only proximal structural change to be a missing lacA 761 gene. This lacA degeneracy may be indicative of its relative functional unimportance 762 compared to the other lac system members 109. 763 Some of the patterns described above can be inferred from coincidence heat charts 764 (Figure 2-4). These matrices represent LMNAST results by the frequency of coincidence 765 between gene characters within hits. The shade of an index represents the frequency of 766 hits where the row gene coincides with the column gene, normalized against the total 767 number of hits containing the row gene, which itself is denoted by (#). For example, in 768 Figure 2-4, the left-most matrix is a representation of a theoretical set of homolog 769 fragments (AB, BC, CD, ABC, BCD, and ABCD). This simple set was constructed to 770 only reflect unbiased homologous recombination presumably resulting only in 771 chromosomal rearrangements. In this set, B and C are extant in five inputs, while A and 772 D are extant in three inputs. All three inputs containing A also contain B, two also 773 contain C, and one also contains D. This is reflected in the shades of the grids in the top 774 row. 775 The middle matrix represents the coincidence distribution amongst LMNAST E. 776 coli lac hits. As an additional example, matrix element (2,1) is a rust color representing 777 the 139 hits with a lacI character of the 155 also containing lacZ. Finally, the right-most 778 matrix is the difference between the left and middle matrices. This particular analysis 779 43 780 Figure 2-4. Coincidence heat map for lac operon LMNAST stringent search hits. Each 781 shaded index represents the normalized frequency of hits containing the row gene that 782 also contain the column gene out of the total number of hits containing the row gene, as 783 denoted by (#). The matrix on the left is a representation of an unbiased set of evenly 784 distributed homologs (AB, BC, CD, ABC, BCD, and ABCD). The middle matrix is the 785 actual coincidence data. The matrix on the right is a heat map of the difference between 786 the two. LacI is heavily over-represented according to the difference matrix. LacY 787 occurred less than would be expected according to a random distribution. 788 789 790 791 792 793 794 44 suggests, for example, that lacI is relatively over-represented across all hits, and that 795 nearly all other coincidences are under-represented; surprisingly, this includes 796 coincidences involving the permease, lacY. Unlike lacA, lacY is believed necessary for 797 lactose catabolism, possibly pointing to the use of a lower affinity transporter in such 798 cases. Obversely, the over representation of lacI indicates a preference for the regulation 799 of lactose catabolism. 800 Of the strong criteria search results, 138 hits contained only two lac gene 801 homologs (average meH 0.28). Two gene homologs represent the natural minimum of 802 individual characters that a homologous system may contain. Such hits represented 803 truncated systems, repurposed individual members, homoplasic convergence, or outright 804 false positives. The majority of these hits fell within clusters of shared Genbank 805 annotation in 2D similarity plots, which compare meHs (averaged BLAST homologies) 806 against LMNAST homologies, or, put differently, average amino acid identities against 807 the system’s broader organizational identity. Generically then, purely vertical 808 displacements imply perfect conservation across species through either vertical or, more 809 likely, recent horizontal gene transfer accompanied by amelioration, while purely 810 horizontal displacements indicate recent gene loss and/or rearrangement. For purposes of 811 downstream analysis, it is interesting to speculate that the kinetics of the remaining genes 812 are unaffected in cases of purely horizontal displacement. For systems subject to HGT, 813 such liberties must necessarily be taken with less confidence. 814 In the case of the stringent lac search, similarity plots revealed a great deal of 815 structural variability in the lac operon homologs of E. coli and near E. coli species 816 (Figure 2-5). Nonetheless, the canonical lac operon (26) and the paralogous evolved 817 45 beta-galactosidase system (43) 110 are clearly the most dominant lac operon-homologs, 818 perhaps partially reflecting the relative preponderance of fully sequenced E. coli strains. 819 Addressing the full breadth of two character homologs, 87 contained lacZ and 820 lacY character types, all of which were adjacent, five of which were misdirected relative 821 to one another. Numerous truncated systems had high meH but imperfect organizational 822 similarity. This cohort was restricted to strains of E. coli and closely related Shigella, 823 Citrobacter, and Enterobacter species, reflecting a generally confined phylogenetic 824 breadth among LMNAST lac hits (Figure 2-3b), and reinforcing the idea of limited lac 825 horizontal gene transfer (HGT) 111. The remainder of the hits consisted of adjacent 826 repurposed characters with functional valence around sugar metabolism. 827 This survey showed that LMNAST E. coli lac operon searches identified 828 numerous ortholog and paralog instances. Relative disparities in gene preservation, gene 829 loss, and structural rearrangements bearing signaling implications were delineated. 830 While there was a significant degree of conformity to the standard genomic arrangement, 831 the amount of diversity indicates that attention paid to related, non-canonical signaling 832 units may be worthwhile. 833 834 46 835 Figure 2-5. 2D similarity plot of lac operon LMNAST stringent search hits overlaid 836 with attributed annotation. Each gray dot represents the homology coordinate of a hit. 837 The size of the dot scales directly with the number of hits at the same coordinate. The 838 dashed line is a 1:1 line along which hits have the same degree of homology by both 839 BLAST and LMNAST measures. Seemingly vertical displacements may imply 840 horizontal gene transfer, while horizontal displacements may imply gene loss or 841 arrangement within the same or proximal species. Ovals indicate a clustering of similarly 842 annotated hits. Dashed ovals denote cases where only the majority of hits therein share 843 47 the labeled Genbank annotation. Here, the dominant features are the original structure 844 and the evolved beta-galactosidase system. Very little HGT is apparent while gene loss 845 and rearrangement are ostensibly more common. 846 847 848 849 850 851 852 853 854 855 856 857 858 859 860 861 48 2.4.2 E. coli K-12 W3110 Lsr System Query 862 Further testing of LMNAST was conducted with weak, stringent, and expanded 863 window searches of the E. coli Lsr system. The query Lsr system consists of a kinase 864 (LsrK: BAA15191), a repressor (LsrR: BAA15192), ABC transporter genes (LsrA: 865 BAA15200, LsrC: BAA15201, LsrD: BAA15202, and LsrB: BAE76456), and AI-2-P 866 processing genes (LsrF: BAE76457, LsrG: BAE76458). Along with AI-2, the Lsr system 867 consists of multiple overlapping positive and negative feedback loops. Multimeric LsrR 868 represses system expression emanating from the intergenic region. AI-2-P, itself 869 catabolized by LsrF and LsrG, allosterically relieves that repression. Thus, both 870 expression troughs and peaks are tightly regulated.112 For the LMNAST search we used 871 the Lsr genes spanning bp 1600331 to 1609003 of E. coli K12 substrain W3110 872 (Genbank nucleotide record AP009048). The number of hits returned using stringent 873 criteria totaled 419. 874 Much like the lac operon, the Lsr system appeared subject to imperfect 875 conservation. Certainly, many fully sequenced E. coli bore exact Lsr homologs (meH > 876 0.95). Exceptions were the truncated systems found in strains BL21 113, REL606 113, and 877 E24377A, and the specific excision of Lsr systems from an otherwise preserved gene 878 order in B2 type E. coli (Figures 2-6 and 2-S1) as revealed through expanded window 879 searching. 880 881 882 883 49 884 Figure 2-6. Lsr system LMNAST hits overlaid onto phylogenetic distributions of 885 different scopes. A E. coli Lsr system LMNAST hits overlaid onto an E. coli specific 886 phylogenetic distribution as developed in and adapted from 102. The larger, bolded leaves 887 contain Lsr system homologs, whereas the grayed italicized leaves do not. Lsr system 888 loss is evident in B2 strains. B E. coli Lsr system LMNAST hits overlaid onto an 889 Enterobacteriales and Pasteurellaceaes phylogenetic distribution adapted from 108. The 890 larger, bolded leaves contain Lsr system homologs, whereas the grayed italicized leaves 891 do not. Loss and gain events are denoted by – and + respectively based on parsimony. 892 Compared to the distribution of the lac operon, Lsr is more phylogenetically dispersed, 893 but also a bit shallower. 894 895 896 50 Unlike the lac operon, numerous Lsr system homologs had perfect LMNAST 897 homology but markedly reduced meH (Figure 2-8a). This is suggestive of amelioration 898 following recent HGT events (which may itself be a reflection of a carefully tuned signal 899 requiring the full complement and correct arrangement of Lsr elements). Indeed, Lsr 900 system GC content varied in accordance with the background GC content, ranging from 901 0.35 to 0.71. Finer scale GC analysis revealed a single consistent and curious feature 902 across all hits with meH greater than 0.3: a sharply spiking dip in fractional GC content 903 near the intergenic region (Figure 2-S2). This dip is suggestive of a conserved DNA 904 binding domain essential to the signal transduction process, which would also, however, 905 be a regulatory feature outside the scope of LMNAST searches. 906 Imperfect LMNAST hits with meH greater than 0.3, deviated from the theoretical 907 distribution according to a bias toward the conservation of lsrB, F, and G, relative to the 908 lsrA, C, and D importer genes (Figure 2-7). This may be attributable to the fact that lsrB, 909 F, and G likely pass cell signaling information downstream 73,74,97 , whereas loss of Lsr 910 importer function might be partially redundant to a low affinity rbs pathway 61, the likely 911 alternate AI-2 import pathway 49,55,112. 912 In contrast to high meH systems, many systems with low meH (less than 0.3) 913 were involved in the metabolism of 5 carbon sugars, mainly ribitol and xylose, according 914 to Genbank annotation (Figure 2-8b). Since AI-2 itself is mainly comprised of a 5 carbon 915 ring, such homology is simultaneously intriguing and unsurprising. More generally 916 among these low similarity hits, lsrK characters were commonly coincident with Lsr 917 51 918 Figure 2-7. Annotated 2D similarity plot for Lsr system LMNAST weak search hits. A. 919 HGT of homologous systems is evident among hits with perfect organizational homology 920 but diminished mean element homology. A great number of hits have low similarity 921 52 along both axes. As seen in B., these hits are mostly involved in the metabolism of 5 922 carbon carbohydrates according to their annotation. This is likely reflective of the fact 923 that AI-2 is of similar structure to 5 carbon sugars. 924 925 926 927 928 929 930 931 932 933 934 935 936 937 938 53 importer characters (lsrA, C, D, and B), indicative of the functional link between these 939 characters. These various features were laid more strongly in relief when measured 940 against the proximal genetic background in an extended window search. 941 While a representation of hit variability preserving structural information can be 942 had from trackback plots (Figure 2-S1), additional salient results from stringent Lsr 943 extended window searching could be evinced from the more summary coincidence heat 944 maps (Figure 2-9). The matrices indicate that lsrK and lsrA genes were strongly 945 preserved among extended window hits. Also, if either lsrF or lsrG were present, the 946 remaining Lsr genes were likely present. The complete system rescission mentioned 947 before was hinted at, especially in rows 3 and 4, corresponding to the toxin/antitoxin 948 hipAB system. Intra-species variation of structural homology increased greatly when 949 using stringent rather than weak criteria (data not shown), mainly as a result of gene loss 950 to pseudo gene conversion, mostly among transporter genes—a bias most easily 951 explained as a matter of pure probability since there are more transporter genes than any 952 other type, and a fact whose functional significance is blunted by an alternative AI-2 953 import pathway. 954 These initial E. coli searches motivated other orthologous Lsr system queries. Full 955 results for E. coli, S. Typhimurium, and B. cereus searches are available in Table 2-S2. 956 These additional searches identified other possible Lsr system homologs, HGT partners, 957 and non-canonical system-associated gene candidates. In Table 2-S2, we delineate 958 operon directionality and gene homology. It is interesting to note that system variants 959 exist among noted human pathogens: Yersinia pestis, Bacillus anthracis, and 960 54 961 962 Figure 2-8. Coincidence matrix for E. coli Lsr system LMNAST stringent search hits. 963 This coincidence matrix depicts the subset of hits with a mean element homology > 0.3 964 and also containing 4-6 gene characters. This subset was chosen for its intermediate 965 degree of homology to the query Lsr system. LsrF and lsrG characters were found to be 966 overrepresented among these hits coincident to hits also containing lsrR and lsrK 967 characters. 968 969 970 971 972 973 974 55 975 Figure 2-9. Coincidence matrix for E. coli Lsr system LMNAST extended window 976 stringent search hits. Letters represent the respective lsr genes. 1-5 represent the five 977 genes preceding the Lsr system: ydeT, yneL, hipA, hipB, and ydeK-lipoprotein. 14-17 978 represent the four genes after the Lsr system: tam (transaconitate methyltransferase), 979 yneE, uxaB, and a predicted diguanylate cyclase. The figure suggests a) strong 980 conservation of the association between Lsr system genes relative to its neighbors, b) hits 981 in which the Lsr system has been excised entirely from its gene neighbors, and c) a weak 982 coincidence between yneH-glutaminase characters and the Lsr system. The matrix also 983 indicates that the overall prevalence of lsrF and lsrG characters is lower than other 984 canonical Lsr characters, although the presence of lsrF and lsrG characters is a good 985 predictor of the presence of other Lsr genes. 986 987 988 56 Haemophilus influenzae. In some instances, lsrRK are either absent (e.g. E. coli BL21) 989 or are associated with altered intergenic regions suggesting altered regulatory control 990 (e.g. Yersinia pestis Antiqua). In other cases transporter genes are distributed with altered 991 bias due to position in the bidirectional operons (e.g. Yersinia pseudotuberculosis 992 PB1/+). In some cases there is no LsrFG component (e.g. Shigella flexerni 2002017). 993 LsrF and LsrG are AI-2-P processing enzymes that lower the cytoplasmic AI-2-P level, 994 thereby contributing to the repression of AI-2 induced genes. 995 Given even only this modest degree of dispersion, it is nonetheless reasonable to 996 suggest that the Lsr autoinduction system is, in fact, extant among scores of bacterial 997 species and that because the organization of genes within the regulatory architecture is 998 varied, the downstream phenotypic behaviors aligned with the AI-2 regulated QS genes is 999 likewise variable. Thus, our results are in line with a general hypothesis that the AI-2 1000 quorum sensing system is broadly distributed and that the specific needs of the bacteria in 1001 a given niche are met by disparate operon arrangements. 1002 The overall phylogenetic distribution of the Lsr system mirrors that as 1003 developed by Pereira, et al. in the cluster they denote as Group I.13 Here, however, 1004 details were fleshed out with different emphases. The Lsr LMNAST search captured the 1005 diversity of pseudo gene conversion, structural rearrangement, and additional hitchhiking 1006 genes associated with the Lsr system that exist in the present nr/nt database. Moreover, 1007 inferences regarding regulatory Lsr system signals could be made that might also map to 1008 phylogenies or possibly, with much more effort, related ecological niches. 1009 1010 57 2.4.3 Analysis of Lsr System Search Results 1011 Results from the various LMNAST searches were reconciled by taking the 1012 highest scoring hit among overlaps within each nucleotide record. Lsr system homologs 1013 clustered mainly in gammaproteobacteria with the greatest density being among E. coli 1014 strains. Diffusely manifesting in more distantly related bacterial species, the Lsr system 1015 appears to have been subject to several HGT events. That is, the Lsr system is absent in 1016 numerous Enterobacteriaceae species, while HGT gain events happened at the root of the 1017 Bacillus cereus group, to R. sphaeroides and R. capsulatus separately, to Sinorhizobium 1018 meliloti, and to Spirochaeta smaragdinae (Figure 2-10). Curiously, while these bacteria 1019 occupy distinct ecological niches, they are all common to soil or water environments. 1020 Multiple extended window searches indicated that S. enterica was the most 1021 proximal cluster for every Lsr system HGT candidate. The sharing of a novel Lsr 1022 system-associated “mannose-6-phosphate isomerase” (NP_460428) between Bacillus 1023 cereus group members, S. smaragdine, and S. enterica, further strengthened the 1024 suggestion of HGT partnership. The gene annotated as “mannose-6-phosphate 1025 isomerase” or “sugar phosphate isomerase,” has recently been shown to be part of the 1026 LsrR regulon in Salmonella.114 Although not part of the E. coli regulon, it was also 1027 associated with S. smaragdinae and B. cereus group orthologs. In keeping with a 1028 possible AI-2-P processing role, it was consistently adjacent to lsrK. 1029 Among gammaproteobacteria, parsimony suggests that two gain events of the Lsr 1030 system occurred: one deeply rooted in enterobacteriales and one in a pasteurellaceaes 1031 ancestor. In the enterobacteriales branch, besides Escherichia, Shigella, and Salmonella, 1032 1033 58 1034 Figure 2-10. Phylogenetic distribution of Lsr at different phylogenetic scales using 1035 reconciled LMNAST results. E. coli Lsr system LMNAST hits overlaid onto a bacterial 1036 phylogenetic distribution as developed in and adapted from 52. Each leaf bears a 1037 representative member from a larger unseen collapsed branch. The larger, bolded leaves 1038 contain Lsr system homologs within the collapsed branch, whereas the grayed italicized 1039 leaves do not. Parenthesized numbers indicate the number of strains and species with Lsr 1040 system LMNAST hits contained within the collapsed branch. 1041 1042 1043 1044 1045 59 Lsr organizational homologs were found in Enterobacter, Photorhabdus, and 1046 Xenorhadbus species, although most of these instances lacked importer genes (lsrACDB). 1047 While it is thought that regulatory proteins conserved across such long 1048 phylogenetic distances often regulate different targets 115, the regulation of community-1049 related functions by different manifestations of the Lsr system (such as biofilm 1050 maturation checkpoints in E. coli 60 and possible biofilm dispersion in B. cereus 82) 1051 suggests a convergent tendency to leverage a quorum/environment sensing capacity 1052 inherent to the Lsr system. Indirect influence over a broader regulon may be abetted by 1053 the involvement of AI-2, the Lsr system substrate, in metabolic pathways 51. 1054 2.4.3.1 Putative Lsr system in Rhizobiales 1055 A grouping of Rhizobiales common to plant symbioses had a conserved set of 1056 adjacent, low homology elements and nonscoring elements that consisted of 1057 unidirectionally expressed genes annotated as: ABC-type transporters (ribose, putative, 1058 putative), a DeoR family repressor (LsrR synonym), aldo-keto reductase (non-scoring), 1059 glycerol-3-phosphate dehydrogenase (nonscoring LsrF synonym), fructose-biphosphate 1060 aldolase (LsrF synonym), and xylulose kinase (LsrK synonym). Somewhat inescapably, 1061 synonymy is at least partly a function of precedence resulting in attribution bias. The 1062 indeterminacy of signaling similarity between these homologs and better characterized 1063 Lsr systems suggests a need for further research. 1064 1065 1066 60 2.5 Concluding Remarks 1067 LMNAST is a program that evaluates similarity or homology on the level of gene 1068 organization, conducting a search for patterns and prevalence constrained by a BLAST E-1069 value filter. Program results overlaid onto phylogenetic data allow visual inspection of 1070 phylogenetic density and dispersion. 2D homology plots display system variability 1071 among LMNAST orthologs, and when overlaid with genera/species clustering, reveal the 1072 degree of system conservation within and across genera/species when organizational 1073 homology decreases and element homology is constant. Clustering also enables the 1074 identification of conserved system homologs. Organizational information is lost when 1075 using coincidence heat charts, but suggestions of the underlying structural variability 1076 remain nonetheless. This is particularly true for coincidence representations of extended 1077 window searches. For these searches, contextual associations with non-canonical genes 1078 also may emerge. Trackback plots illustrate both variety and structural information, 1079 albeit in a less dense format. These representations are especially useful in combination. 1080 It should be noted that the results are nearly entirely comprised of excerpts from fully 1081 sequenced genomes. Results are also biased by BLAST input, as characters with more 1082 element homologs (e.g. lsrA) appear more frequently in hits. 1083 Generically, LMNAST identified query homologs with a variety of deletions, 1084 insertions, misordering, and misdirections. While nearly any source of mutagenesis may 1085 result in a frank mutation, affecting a system’s organizational homology, homologous 1086 recombination, insertion sequences, transposable elements, and combinations thereof are 1087 likely to be of particular consequence for LMNAST searches. Deletions may be a result 1088 of pseudo gene conversion, of chromosomal rearrangements, or part and parcel of an 1089 61 insertion event—if the insertion results in a gap sufficiently large as to disconnect hit 1090 elements from one another. In the case of such insertions, sufficiently weak criteria may 1091 be of use, with the caveat that decreased stringency increases the number of false 1092 positives. From a signaling perspective, depending on the impacted elements and the 1093 nature of the inserted sequence, gap presence could result in system discoordination; and 1094 the longer the gap the more probable and severe the discoordination, most likely to the 1095 detriment of system function. 1096 As for the specific test queries examined herein, while the lac operon is well 1097 characterized in its canonical form, there nonetheless exists a great deal of frank variation 1098 from the textbook case. Of particular interest are homologous instances where structural 1099 rearrangement could influence self-regulation of component expression. Also of note are 1100 its multiple signaling component deletions. Such abbreviated modules are frequently 1101 repurposed in a related context. Complete lac operons were found among nearly all E. 1102 coli strains. Incomplete lac operons were found to be distributed only among closely 1103 related Enterobacteriaceae species comprised almost entirely of Escherichia, 1104 Citrobacters, Enterobacters, and Serratias as expected based on limited lac operon HGT 1105 111. This difference between the rates of decay for the two homology signals over 1106 phylogenetic space may be suggestive of distinct selection pressures guiding the two 1107 systems. Also identified through LMNAST were conserved, E. coli-specific evolved 1108 beta-galactosidase systems 110, demonstrating a capacity to find closely related systems. 1109 On par, Lsr system hit structural similarity was less well correlated with meH than 1110 lac operon results, a phenomenon presumably associated with apparent Lsr system HGT. 1111 The Lsr system was phylogenetically dispersed more widely than the lac operon, even 1112 62 while its distribution remained densest among gammaproteobacteria. Much like the lac 1113 operon, Lsr system structure was subject to significant variability. lsrK and lsrR 1114 characters were common to many hits. lsrF and lsrG were the least common; the 1115 inclusion of both elements nearly always coincided with the presence of all other Lsr 1116 characters as well. Lsr-contextually associated genes and novel putative Lsr systems 1117 were also elucidated. 1118 The dispersion of Lsr to bacteria as far afield as the S. smaragdinae, the first 1119 Spirochaeta to be fully sequenced 116, is intriguing. It suggests that while the depth of 1120 Lsr dispersion may not be significant, that its exposed breadth will expand incrementally 1121 at a rate proportional to microbial genome sequencing. While the direct regulon of such 1122 HGT systems is expected to be limited 69,115, the proximity of the substrate to key 1123 metabolic pathways may allow the Lsr system to confer contextual phenotypic 1124 advantages by impacting downstream pathways with its capacity to recompartmentalize a 1125 metabolic intermediate. Moreover, the known regulatory requirements for functional 1126 integration of the Lsr system are minimal, consisting entirely of interaction with cAMP-1127 CRP complex, which is deeply rooted in eubacteria. Gene organization differences 1128 between dispersed Lsr homologs, may indicate distinct signaling outcomes, in turn 1129 suggesting the appropriation of the Lsr system’s inherent quorum capacity to drive 1130 distinct phenotypes suited to a given bacteria’s needs within its particular niche. 1131 Unlike the results for the lac operon, Lsr system results returned a large number 1132 of other-annotated, low homology systems. This speaks to both the inherent difficulty of 1133 extrapolating based on homology and the utility of the additional, complementary 1134 homology measure yielded by LMNAST searching. Overall, given the complexity of the 1135 63 results, numerous aspects may be of interest. Some graphical tools of a complementary 1136 nature (e.g. 2D similarity plots and coincidence heat maps) have been used here for 1137 distillation and closer inspection. The extant variation of the queried modular systems, as 1138 captured by frank changes in gene organization, was revealed. 1139 1140 1141 1142 1143 1144 1145 1146 1147 1148 1149 1150 1151 1152 1153 64 2.6 Supplemental Materials 1154 1155 Figure 2-S1. Trackback plots for Lsr system LMNAST extended window stringent 1156 search hits. These diagrams describe the variety of LMNAST hits in greater detail. A 1157 straight diagonal line indicates complete agreement with the query. Rearrangements are 1158 represented by discontinuities. Relative redirection is indicated by a flipping of the 1159 diagonal orientation. Deletion is indicated by horizontal dashed gaps. Insertion is 1160 indicated by vertical gaps. The legend in the upper right hand corner indicates which 1161 numbers correspond to which genes. Trackback plots are organized into categories: A, B, 1162 C, D, and E according to the following categories: I. Prototype Lsr systems, II. Modified 1163 Lsr systems with pre and post-Lsr adjacent characters, III. Modified Lsr systems with 1164 post-Lsr adjacent characters, IV. Modified systems without Lsr adjacent characters, and 1165 V. Highly modified Lsr systems. For exact subgroup membership see Table S1. 1166 1167 1168 65 1169 Table 2-S1 1170 Group Subgr oup Species-Strain I (pre-Lsr-post) A Escherichia coli ATCC 8739 Escherichia coli DH1 B Escherichia coli O103:H2 str. 12009 Escherichia coli O157:H7 str. EC4115 Escherichia coli SE11 C Escherichia coli O111:H- str. 11128 Escherichia coli O157:H7 str. TW14359 Escherichia coli O26:H11 str. 11368 Escherichia coli 55989 chromosome D Escherichia coli IAI1 E Escherichia coli BW2952 Escherichia coli str. K12 substr. DH10B F Escherichia coli O157:H7 str. Sakai Escherichia coli O157:H7 EDL933 G Escherichia coli SMS-3-5 II (pre-modified Lsr- post) A Escherichia coli E24377A B Escherichia coli B str. REL606 Escherichia coli BL21(DE3) C Escherichia coli IAI39 D Shigella dysenteriae Sd197 E Escherichia coli 536 Escherichia coli APEC O1 Escherichia coli CFT073 Escherichia coli IHE3034 Escherichia coli S88 chromosome Escherichia coli SE15 Escherichia coli UTI89 F Escherichia coli ED1a chromosome G Escherichia coli 042 Shigella boydii CDC 3083-94 III (modified Lsr-post) A Escherichia fergusonii ATCC 35469 B Shigella flexneri 2a str. 2457T Shigella flexneri 2a str. 301 C Sinorhizobium meliloti 1021 plasmid pSymB Rhodobacter sphaeroides KD131 chromosome 2 IV (modified Lsr) A Aggregatibacter aphrophilus NJ8700, complete genome Enterobacter sp. 638 Haemophilus influenzae PittEE Haemophilus somnus 129PT Haemophilus somnus 2336 66 Klebsiella pneumoniae 342 Klebsiella pneumoniae subsp. pneumoniae MGH 78578 Pasteurella multocida subsp. multocida str. Pm70 Salmonella enterica subsp. enterica serovar Agona str. SL483 Salmonella enterica subsp. enterica serovar Choleraesuis str. SC-B67 Salmonella enterica subsp. enterica serovar Enteritidis str. P125109 Salmonella enterica subsp. enterica serovar Gallinarum str. 287/91 Salmonella enterica subsp. enterica serovar Heidelberg str. SL476 Salmonella enterica subsp. enterica serovar Paratyphi A str. AKU_12601 Salmonella enterica subsp. enterica serovar Paratyphi A str. ATCC 9150 Salmonella enterica subsp. enterica serovar Paratyphi B str. SPB7 Salmonella enterica subsp. enterica serovar Paratyphi C str. RKS 4594 Salmonella enterica subsp. enterica serovar Schwarzengrund str. CVM19633 Salmonella enterica subsp. enterica serovar Typhi Ty2 Salmonella enterica subsp. enterica serovar Typhimurium str. D23580 Salmonella enterica subsp. enterica serovar Typhimurium str. LT2 Yersinia enterocolitica subsp. enterocolitica 8081 Yersinia pestis Angola Yersinia pestis CO92 Yersinia pseudotuberculosis IP 31758 Yersinia pseudotuberculosis IP32953 Yersinia pseudotuberculosis PB1/+ Yersinia pseudotuberculosis YPIII B Klebsiella pneumoniae NTUH-K2044 DNA Salmonella enterica subsp. enterica serovar Dublin str. CT_02021853 Salmonella enterica subsp. enterica serovar Newport str. SL254 Salmonella enterica subsp. enterica serovar Typhimurium str. 14082S Yersinia pestis biovar Microtus str. 91001 Yersinia pestis Pestoides F C Yersinia pestis Antiqua Yersinia pestis Nepal516 D Yersinia pestis KIM E Rhodobacter sphaeroides 2.4.1 chromosome 2 Rhodobacter sphaeroides ATCC 17029 chromosome 2 F Bacillus anthracis str. ‘Ames Ancestor’ Bacillus anthracis str. Ames G Bacillus anthracis str. A0248 H Bacillus anthracis str. CDC 684 Bacillus anthracis str. Sterne Bacillus cereus ATCC 10987 Bacillus cereus ATCC 14579 Bacillus cereus B4264 Bacillus cereus E33L Bacillus cereus G9842 67 Bacillus cereus Q1 Bacillus thuringiensis serovar konkukian str. 97-27 Bacillus thuringiensis str. Al Hakam Bacillus weihenstephanensis KBAB4 Shigella sonnei Ss046 I Bacillus cereus AH187 Bacillus cereus AH820 Bacillus cereus 03BB102 V (Non-continous) A Escherichia coli ED1a chromosome Serratia proteamaculans 568 Shewanella halifaxensis HAW-EB4 Yersinia pestis Angola Yersinia pestis Antiqua Yersinia pestis biovar Microtus str. 91001 Yersinia pestis CO92 Yersinia pestis KIM Yersinia pestis Nepal516 Yersinia pestis Pestoides F Yersinia pseudotuberculosis IP31758 Yersinia pseudotuberculosis IP32953 Yersinia pseudotuberculosis PB1/+ Yersinia pseudotuberculosis YPIII B Agrobacterium radiobacter K84 chromosome 2 Ochrobactrum anthropi ATCC 49188 chromosome 2 Rhizobium leguminosarum bv. trifolii WSM1325 plasmid pR132501 Rhizobium leguminosarum bv. viciae plasmid pRL12 Rhizobium sp. NGR234 plasmid pNGR234b Silicibacter sp. TM1040 Sinorhizobium medicae WSM419 plasmid pSMED01 Sinorhizobium meliloti 1021 plasmid pSymB C Burkholderia cenocepacia AU 1054 chromosome 1 Burkholderia cenocepacia HI2424 chromosome 1 Burkholderia cenocepacia J2315 chromosome 1 Burkholderia cenocepacia MC0-3 chromosome 1 Burkholderia multivorans ATCC 17616 DNA Burkholderia sp. 383 chromosome 1 Burkholderia xenovorans LB400 chromosome 1 1171 1172 1173 68 1174 Figure 2-S2. GC content demonstrates consistent spiking dip at intergenic region. GC 1175 content graphs for Bacillus cereus ATCC10967, Yersinia Pestis Nepal516, Rhodobacter 1176 sphaeroides ATCC 17029, and Yersinia pestis CO92. Graphs are labeled with lsr system 1177 beginning (end of lsrK), lsrR gene intersection with the intergenic region (labeled lsrR 1178 end), and lsr system ending (end of lsrG). Arrow direction indicates the direction of 1179 lsrACDBFG expression. Proximal to the intergenic region is a conserved dip in GC 1180 content. 1181 1182 1183 1184 69 1185 70 1186 71 1187 72 1188 73 Table 2-S2. Results from three separate LMNAST searches for Lsr system homologs. 1189 The E. coli K-12 W3110 search shown was completed using weak criteria (a lower gap 1190 extension penalty and less rigid adherence to annotation), whereas the B. Cereus ATCC 1191 10987 and S. enterica Typhimurium LT2 searches used stringent criteria (higher penalties 1192 for deviation from the query pattern and adherence to supplied annotation). “Join” 1193 indicates a difficulty in handling irregular annotation where the first gene annotated in a 1194 record spans the end and the beginning of the record. Arrow direction indicates the 1195 direction of transcription along the genome. Color is a stand-in for character type, and 1196 the degree of shading indicates degree of element homology, with the darkest shade 1197 representing 100% element homology. 1198 1199 1200 1201 1202 1203 1204 1205 1206 1207 74 Chapter 3: Quorum desynchronization leads to bimodality and 1208 patterned behaviors in microbial consortia 1209 3.1 Abstract 1210 Quorum sensing (QS) is a type of signaling used to coordinate behavior in bacterial 1211 consortia via secreted autoinducers. Here we focus on Lsr (LuxS regulated) based QS 1212 signaling that activates only a fraction of a cell population through the ‘universal’ QS 1213 signal autoinducer-2 (AI-2). Our modeling indicates that bimodality arises from 1214 desynchronized, QS-activated importation of AI-2, and is sensitive to cell-cell distance 1215 and cell density. According to our agent-based models, through this mechanism, we 1216 found Lsr QS drives spatial organization of cell signaling. That is, Lsr induced AI-2 1217 internalization results in emergent “cluster-then-disperse” behavior when acting in 1218 concert with AI-2 chemoattraction, and speckled activation in surface attached bacterial 1219 colonies. This is contrasted against rapid and population-wide expansion of QS activity 1220 emerging from LuxIR based QS in a growing bacterial colony. Further, Lsr signaling 1221 also results in differential QS activation between Lsr/LuxS types in in silico mixed 1222 cultures. One particular finding was that signal-negative cells were also found to be 1223 effectively signal-blind in Lsr QS. More broadly, interstrain communication modulated 1224 QS activation patterns were elucidated, helping to frame the complex Lsr QS dynamics 1225 arising in sociomicrobiological settings. 1226 1227 1228 75 3.2 Introduction 1229 Quorum sensing (QS) is a bacterial response to self-secreted signaling molecules 1230 generically known as “autoinducers”. While QS has been observed among individual 1231 bacteria in experimentally manipulated settings 4–6, QS is often described as informing 1232 the coordination of processes (such as virulence factor production and biofilm formation) 1233 that are metabolically burdensome and ineffectual for individual cells, yet beneficial at 1234 multicellular or population scales 80. Population scale coordination typically arises from 1235 the decompartmentalization of the autoinducer that acts as a shared pool of extracellular 1236 signal available to coordinate individual cells that are sufficiently proximal. This 1237 regulatory strategy can reduce “noisy” inputs and other heterogeneity by focusing 1238 phenotypic outcomes and organizing population activity 117,118. Coordination varies in 1239 degree and fashion, and depends upon the mechanistic underpinnings of the QS system 1240 involved 94,119,120. Importantly, a given “quorum’s” collective response to heterogeneity 1241 is likely to reflect the topology of the cognate QS signaling “module” 121,122. Among the 1242 better studied QS systems are LuxIR (background Figure 3-1), which is widely utilized 1243 in synthetic biology because of its genetic simplicity, and Lsr (foreground Figure 3-1), 1244 which is widely distributed among gammproteobacteria 14. LuxIR is induced by 1245 autoinducer-1 (AI-1) (an umbrella term encompassing a variety of species specific acyl-1246 homoserine lactones (AHLs)) whereas Lsr is induced by autoinducer-2 (AI-2) (which is 1247 produced by luxS in nearly half of eubacteria 46). LuxIR signaling is comprised of an 1248 intracellular positive feedback loop 21 that is tied to positive intercellular feedback at the 1249 population scale. This is contrasted against Lsr signaling the induction of which features 1250 76 1251 Figure 3-1. LuxIR and Lsr QS activity intracellularly and intercellularly. The LuxIR 1252 system (background) involves a simple intercellular feedforward loop that entangles AHL 1253 synthase production with LuxR regulatory production, which itself promotes the 1254 expression of both LuxI and LuxR when AHL concentration is sufficiently high. This 1255 loop is connected to intercellular feedforward activity due to passive diffusion of AHL 1256 across the cell membrane. This is in contrast to the Lsr system (foreground), where the 1257 feedforward loop of LsrACDB increasing AI-2 intracellularly, LsrK phosphorylating AI-1258 2, and AI2-P derepressing LsrR which in turn stimulates LsrACDB and LsrK expression 1259 does not drive extracellular AI-2 concentrations higher, but instead depletes them, 1260 creating negative intercellular feedback. 1261 1262 1263 1264 1265 1266 1267 1268 77 the rapid depletion of extracellular AI-2 55, intertwining intracellular positive feedback 1269 with extracellular negative feedback. Lsr activity is subject to further negative feedback 1270 by the enzymatic processing of phosphorylated AI-2 through LsrF and LsrG 73–75. Among 1271 the small molecules produced by LsrF and LsrG under aerobic conditions are glycerol-3-1272 phosphate 123 and phosphoglycolic acid 75, respectively, which serve as metabolic 1273 intermediates for AI-2 reassimilation into primary metabolism. 1274 As described herein, we explored and contrasted QS operation in a variety of 1275 contexts through the use of finite element-agent based models. In part, our modeling 1276 efforts were motivated by indications that bimodal Lsr expression may arise in pure 1277 cultures, where cell populations developed into QS-activated and non-QS activated 1278 fractions 94. While it has been suggested that a transient bimodality may arise as a 1279 consequence of intracellular signaling topology 95, we hypothesized that a more 1280 permanent population bifurcation might develop from hyperlocal competition for AI-2 1281 between cells. This possibility was probed here by employing a population of ODEs 1282 wherein nongenetic heterogeneity was explicitly incorporated. The ramifications of such 1283 a hyperlocal competition were then considered in the contexts of cell motility and mixed 1284 consortia. 1285 LuxIR and Lsr QS were contrasted and found to generate different spatial patterns 1286 of cell signaling in silico. We recapitulated an earlier experimental study wherein the 1287 sudden expansion of LuxIR QS activation from a cell colony center (or “Supernova”) 1288 was observed 33. This stood in clear contrast to speckled QS activation found in Lsr-1289 mediated cell colonies. Further, the confluence of Lsr QS and AI-2 chemoattraction in E. 1290 coli 97 was examined, the combination of which produced a “cluster-then-disperse” 1291 78 behavior that is consistent with concepts of sociomicrobiology 124 wherein “travel” to 1292 better locales may be one of the currencies of public goods 125. 1293 Additionally, the consequences of Lsr QS circuitry were examined in silico in 1294 mixed cultures of different Lsr/LuxS cell types. luxS and lsrFG mutants were chosen 1295 for their prevalence and because they represent slower and faster QS activating 1296 populations, respectively. Coupling these mutants with their wildtype counterparts 1297 illustrated how Lsr dynamics manifest in the preferential activation of one Lsr/LuxS cell 1298 population over another. Just as activation patterns arising from Lsr signaling were distinct 1299 from those associated with LuxIR activity, interpopulation transactions between 1300 Lsr/LuxS subgroups revealed different incentives for social cheating than have previously 1301 been associated with generic QS signaling. 1302 1303 1304 1305 1306 1307 1308 1309 1310 1311 79 Methods 3.3 1312 3.3.1 Modeled Cell Behaviors 1313 During a given time step, based on cell growth characteristics and local 1314 concentration of substrate, cells divided, moved, and responded to autoinducer 1315 concentrations in their direct extracellular environment through an inflection of their QS 1316 dynamics as modeled by ODE trajectories. AI-2 was also exchanged between grids in an 1317 approximation of diffusion. Full model details and in particular those common to all 1318 simulations are provided in SI Methods. 1319 3.3.1.1 Chemotactic Swimming. To interrogate QS activation patterning under different 1320 regimes of motion, the mode of cell motility was varied. To begin either swimming or 1321 colony growth simulations, cells or colony centers were placed randomly in the simulated 1322 environment. In simulations of chemotactic behavior, cells adjusted direction every 0.1 1323 seconds. If the average concentration of AI-2 experienced by the cell over the previous 1324 0.1 seconds was less than a minimum sensitivity or less than the average concentration of 1325 AI-2 from the previous 15 second period (reflecting the time required for accommodation 1326 126), the cell moved in a random direction and at a speed produced from a distribution 1327 with an average of 20 m/sec, in accordance with results from cell tracking studies 1328 [unpublished]. If the short term average concentration was higher than the long term 1329 average, the cell continued moving in the same direction as it had the previous time step, 1330 with slight error. 1331 3.3.1.2 Colony Growth. Expanding colony growth was loosely mimicked by cells 1332 moving according to space filling considerations. In the interior of the colony, daughter 1333 80 cells pushed neighbors outward based on the shortest distance to the colony boundary 1334 available. At colony boundaries, cells moved into randomly chosen free space. All 1335 motions were constrained by simulation boundaries. 1336 3.3.1.3 LuxIR QS. To simulate LuxIR behavior, cells produced AHL at high, baseline, or 1337 intermediate rates (10 M/min to 1 M/min) corresponding to active, inactive, or linearly 1338 activating QS, respectively. The rate of AHL production in the QS inactive state 1339 (baseline AHL production) was varied across the population with the same coefficient of 1340 variation that was applied to the parameter basal in Lsr simulations ( = 0.0225, log-1341 normal distributions). Diffusion across the membrane was modeled at a conductivity of 1342 0.6 between the intracellular and extracellular spaces driven by concentration difference 1343 127. LuxIR expression was activated over a thirty minute period in a linear progression 1344 once an intracellular AHL threshold of 2.9 M was exceeded, coarsely capturing inherent 1345 system cooperativity and the time lag associated with transcription and translation. 1346 3.3.1.4 Lsr QS. The ODEs governing Lsr activity, associated parameters, and the 1347 rationale behind the design choices are described in SI. Lsr related ODEs relied heavily 1348 upon Michaelis-Menten kinetics and cooperative Hill terms. Parameters were chosen to 1349 be biologically realistic and to conform to the behavior of the system as gleaned from 1350 previously conducted experiments (see SI) 55,56,94,112. 1351 3.3.2 Simulation Variants: Gene Deletions and Mixed Populations 1352 As metabolic burdens 128 were not considered, gene deletions were modeled by 1353 setting the appropriate parameter to zero, reflecting ablated activity. The gene deletions 1354 examined consisted of luxS and lsrFG. These were modeled by setting parameter 1355 81 values to zero such that Ksynth = 0 and kCat = 0, respectively. These variants differed in 1356 their sensitivity to external AI-2. Simulations placing variant populations and wildtype 1357 cells in the same environment were run. The two populations were equivalently 1358 populated to initiate the simulation and divided according to the same heuristics. 1359 Chemotaxing motility was stripped from these simulations in order to focus on the effect 1360 from deletion of Lsr and auxiliary proteins. 1361 In each of these cases, regardless of the cell type simulated, all bacteria and 1362 growing colonies were accounted for at every time step and in every grid point. Owing 1363 to the complex nature of the system, a finite difference method was used. Care was taken 1364 to minimize error propagation (studies of grid size, time step, growth parameters, etc.). 1365 1366 1367 1368 1369 1370 1371 1372 1373 1374 82 3.4 Results 1375 3.4.1 Lsr Autoinduction in Pure Cultures. We first modeled desynchronized Lsr QS 1376 activation. Specifically, the value of the model parameter basal (the rate of AI-2 transport 1377 through the low-flux AI-2 importer), was distributed among modeled cells in a log 1378 normal fashion. All other parameters were held constant so that the relative background 1379 AI-2 uptake rate and corresponding speed to Lsr QS activation were the only 1380 distinguishing features among the entire cell population. We observed that only a 1381 fraction of the total cell population became QS activated, generating a bimodal pattern. 1382 The exact on/off balance was influenced by the degree of variance in the basal 1383 distribution (Figure 3-2A). As discussed extensively in the SI, this bimodality ostensibly 1384 resulted from QS active cells depriving inactive cells of AI-2 with which to induce. 1385 Consider an ith cell with basali << basalMean that would undergo QS activation if basali = 1386 basalj for all i,j, instead remaining inactive. When the coefficient of variance was zero, 1387 the entire population became activated almost simultaneously. Increasing the variance of 1388 the value for basal increased the population fraction remaining inactive, attributable to 1389 cells with higher values of basal from the leading edge of the distribution activating 1390 earlier and earlier as the variance of the distribution increased. The earlier the activation 1391 of these cells relative to the mean, the faster extracellular AI-2 was drawn down, thus 1392 preventing the activation of cells with lower basal values, which themselves required 1393 more and more time to accumulate sufficient AI-2P. This conceptual model serves as a 1394 basis to explain previously reported flow cytometry data, where only about half of the 1395 total population took on the QS-activated phenotype 94. 1396 1397 83 Figure 3-2. Lsr autoinduction of 1398 pure cultures leads to bimodal 1399 phenotype. A The fraction of the 1400 population QS activated over time 1401 was influenced by the standard 1402 deviation of the natural logarithm, 1403 , of the log-normal distribution 1404 for the parameter basal, here for  1405 values ranging from 0 to 0.025, run 1406 in triplicate. Dark lines represent 1407 the average values, whereas lighter 1408 surrounding lines represent the 1409 standard deviation. Activation was 1410 the slowest but most abrupt for the 1411 cell population with a variance of 1412 zero, quickly achieving complete 1413 activation. Inset is the distribution 1414 of the value of basal for three 1415 different values of . B Inset is the 1416 plateaued value of the population 1417 fraction that was Lsr active for 1418 populations of different median 1419 basal values. In the main figure, 1420 the overall extracellular 1421 concentration of AI-2 per Lsr 1422 induced cell is represented on the 1423 primary ordinate axis. Each curve 1424 represents an average from three 1425 different simulations of non-taxis 1426 swimming cells. The median 1427 minimal distance between cells 1428 averaged for 20 simulations of non-1429 taxis swimming cells is represented 1430 on the secondary ordinate axis. At 1431 the right top is a pictorial example 1432 of the minimum distance between 1433 cells (where dmin = min (d1, d2, d3,... 1434 dn, n=6). 1435 1436 1437 1438 84 We also note that the distribution of the bifurcation was possibly influenced by 1439 the related but competing factors of intercellular distance and cellular concentration. This 1440 was suggested by a general correlation between decreasing basal and decreasing QS 1441 activation as seen in the inset of Figure 3-2B. As indicated by the timing of spikes in 1442 Figure 3-2B, reducing basal resulted in delayed activation. This delayed activation 1443 corresponded to diminished intercellular distance due to growth (light blue line) 1444 calculated as the minimum distance between neighboring cells (a cartoon of which is 1445 depicted on the top right corner of Figure 3-2B for an ith cell). Increased cell proximity 1446 may have amplified the negative feedback associated with QS activation, as the local 1447 depletion of AI-2 by QS active cells was able to delay or prevent the activation of more 1448 cells. Further evidence of this cell-cell distance effect is provided in the SI. The putative 1449 influence of intercellular distance appeared to be constrained, however. That is, when cell 1450 concentration was too high, more QS activated cells were required to draw down local 1451 AI-2 concentrations as indicated by the relatively broad peak of the last curve depicting 1452 AI2 concentration divided by the number of QS activated cells. Subsequently, depletion 1453 of AI-2 from the environment was delayed, allowing more cells to become QS activated 1454 than would have otherwise. This competing effect was ostensibly dominant for 1455 populations with the lowest median value of basal in Figure 3-2B, as the QS activated 1456 fraction increased instead of continuing to decrease. 1457 3.4.2 Pattern formation in QS systems: LuxIR vs Lsr. In Figure 3-3A, we show that a 1458 “supernova-like” LuxIR QS activation pattern emerged from near the colony center, 1459 loosely recapitulating experimental results seen with LuxIR engineered cells 33,129. For 1460 this simulation, cell motility was limited to gliding along the surface in a space filling 1461 85 capacity purely as a function of cell outgrowth (cell division resulting in occupation of 1462 otherwise empty neighboring site by daughter cell) without directional bias. Once 1463 initiated, LuxIR activation (red color) spread quickly outward from its point of origin 1464 often near the center of colony growth (due to positive feedback processes of cell growth 1465 and autoinduced AHL production), engulfing inactive neighbors in a building wave of 1466 high AHL concentration, coordinating population expression in what appears as a 1467 traveling wave front of activated cells. 1468 In Figure 3-3B, we depict simulated behavior for the Lsr system with identical 1469 gliding motility rules as for the LuxIR simulations. Instead of a blossoming wave of QS 1470 activation (Figure 3-3A), scattered patches of activation appeared, presumably 1471 influenced by the same Lsr driven AI-2 recompartmentalization dynamics associated with 1472 bimodal activation. Undergirding the distinction between these activation patterns were 1473 differences in the interplay between non-genetic heterogeneity and intercellular QS 1474 feedback, as discussed in SI, where activation heterogeneity was also quantified. Thus, 1475 in the case of Lsr, QS activation was not apparently localized, nor was it uniform. 1476 Instead, fractional QS activation was observed among the population whole. 1477 3.3.3 Cell Motility – Lsr QS based pattern emergence. Swimming was modeled either 1478 as an unbiased process or was governed by heuristics approximating AI-2 1479 chemoattraction, described in the methods. Early in the simulations, prior to QS 1480 activation, chemotaxing cells assembled into clusters 130, attracted to each other by AI-2 1481 molecules. Largely, cell populations gradually coalesced along simulation boundaries as 1482 our boundary conditions preclude diffusion through these boundaries. Since clusters 1483 86 Figure 3-3. QS dynamics coupled 1484 with gliding during colony growth. 1485 A Images from a representative 1486 simulation for LuxIR/AHL 1487 dynamics coupled with colony 1488 growth. QS active cells are in red, 1489 whereas inactive cells are in yellow. 1490 B Images from a representative 1491 simulation of Lsr/AI-2 dynamics 1492 coupled with colony growth. QS 1493 active cells are in red, whereas 1494 uninduced cells are yellow. 1495 1496 1497 1498 1499 1500 1501 1502 1503 1504 1505 1506 1507 87 were formations of relatively high local cell density and correspondingly high AI-2 1508 concentration, cells within clusters activated more quickly than unclustered cells. Once 1509 Lsr activity was sufficiently pervasive, clusters were transformed into sites of the most 1510 rapid AI-2 depletion, dissipating the AI-2 gradient and the impetus for clustering, the 1511 ultimate result being cluster dispersal. This emergent behavior, depicted here for the first 1512 time in a model, generates an interesting overall “collect-disperse” pattern. This pattern 1513 is exemplified in Figure 3-4. At zero minutes a randomized distribution of an initial cell 1514 population was observed. At 100 minutes, cell clusters began to coalesce, as seen at the 1515 bottom and left edge and near the top left corner of the environment. At 200 minutes, 1516 clusters were more pronounced, and at 244 minutes, QS activated cells began to emerge 1517 (red). At 266 minutes more cells were QS active and the clusters became more dispersed. 1518 By 300 minutes clusters had become entirely dispersed. Through many repeated 1519 simulations, the exact placement and number of clusters were found to be inconsistent. 1520 Quantification of this clustering (by cell-cell distance) is available in SI. Interestingly, as 1521 different motility modes implied different cell-cell proximity, motility appeared to 1522 feedback onto fractional activation, which is also discussed in SI. 1523 3.4.4 Mixed Population Simulations. Additional consequences of desynchronized Lsr 1524 based AI-2 recompartmentalization were identified through the study of competition for 1525 AI-2 among in silico mixed cultures of a wildtype population with varied derivative 1526 mutant populations. Here, initial conditions were the same as in previous simulations, 1527 except that half the modeled cell population had modulated Lsr/LuxS activity, reflecting 1528 different genotypes. Figure 3-5A represents the results from a wildtype pure culture as a 1529 88 1530 1531 1532 1533 1534 1535 1536 Figure 3-4. Cluster-disperse pattern from combination of Lsr and chemotaxis. Images 1537 from a representative simulation of Lsr/AI-2 dynamics coupled with chemoattraction to 1538 AI-2. Autoinduced cells appear in red, whereas uninduced cells appear in yellow. 1539 1540 1541 1542 1543 1544 1545 1546 1547 1548 1549 1550 89 1551 Figure 3-5. Mixed culture simulations. Each graph may be considered as being 1552 composed of a series of histograms over time. The vertical axis represents frequency, the 1553 right axis (towards the reader) represents increasing log transformed concentration of 1554 transporter protein, and the left hand axis (away from the reader) represents increasing 1555 time in minutes. A wildtype cells alone. BluxS populations mixed with wildtype cells. 1556 C lsrFG cells mixed with wildtype cells. 1557 1558 1559 1560 1561 1562 1563 1564 90 reference (lower panel). Initially, all cells were QS inactive (represented here as 1565 normalized to 0.0). Around 200 minutes, subpopulations began to branch off into high 1566 expressers, representing QS activation in these cells. In Figure 3-5B where luxS 1567 (above) and wildtype (below) cells were cultured together in silico, wildtype cells 1568 became dominantly activated over the course of the simulation, with luxS cells 1569 remaining largely Lsr inactive. In fact, even by the end of the simulation, luxS 1570 population Lsr expression levels were only nominally above their baseline. 1571 As in Figure 3-5C, mixing of wildtype cells (below) with lsrFG cells (above) 1572 also resulted in uneven activation between populations, but for ostensibly different 1573 reasons. Here, lsrFG cells were dominantly activated and became so early on, due to 1574 the early activation of lsrFG cells, which accumulate AI2-P more readily than wildtype 1575 cells 131. This then deprived wildtype cells of AI-2, resulting in a smaller activated 1576 fraction than when wildtype cells were cultured by themselves. 1577 1578 1579 1580 1581 1582 1583 91 3.5 Discussion 1584 Bimodal expression arising from clonal origins is a common phenomenon, 1585 frequently associated with pattern formation and differentiation in multicellular organisms 1586 132,133. As is commonly the case and here also, population bifurcation is expected to arise 1587 as a consequence of nonlinear responses to nongenetic heterogeneity manifested in bimodal 1588 protein expression 134,135. Specifically, our results suggest that bimodal expression arises 1589 when the intercellular negative feedback associated with Lsr activated AI-2 1590 recompartmentalization is desynchronized across a population (here represented by varied 1591 rates of AI-2 uptake and subsequent Lsr mediated gene expression). This bimodal 1592 expression may represent role diversification. Bacterial diversification is frequently framed 1593 within the context of bet hedging 136 or as a graded response to environmental conditions, 1594 an example being the different conditions that at the margins of biofilms compared to those 1595 within the bulk 137. Although Lsr signaling does influence biofilm development in E. coli 1596 60, whether population diversification through Lsr QS represents bet-hedging or some other 1597 transient specialization within the context of a population-wide transition to a sessile 1598 lifestyle remains a point of significant interest. Regardless, in silico experiments here 1599 clearly frame Lsr QS activity within a context of phenotype diversification. We further 1600 explored how this diversification operates spatially and between strains with different 1601 Lsr/LuxS activity. Spatial self-arrangement of Lsr activity was evaluated and contrasted 1602 against that of LuxIR. LuxIR dynamics produced a sudden rapidly expanding activation 1603 from the colony center. With an autoinducer that requires active transport between cell 1604 compartments, the Lsr system provided a stark contrast by producing speckled activation. 1605 This distinction may be attributed to the coupling of positive intracellular and extracellular 1606 92 feedbacks to one another in LuxIR QS, whereas intracellular positive feedback drives an 1607 extracellular negative feedback in Lsr QS (recall Figure 3-1). 1608 Lsr QS populations subject to AI-2 chemoattraction displayed an interesting 1609 collect-disperse behavior, which informed a complete cycle of nutritional resource 1610 exploitation: seek, find, deplete, and seek again. To our best knowledge, this is the first 1611 time that such behavior has been characterized by any mathematical model or otherwise 1612 described in the context of the specific QS mechanisms investigated here. Since 1613 extracellular AI-2 could be viewed as a potential marker for the exponential growth of a 1614 broad swath of bacterial species (its production is a direct consequence of primary carbon 1615 metabolism) 27,46, motility toward such active growth is likely to be most fruitful for a 1616 generalist like E. coli, potentially reinforcing chemotactic tendencies toward other 1617 substrates. As catabolite repression is epistatic to QS switching/Lsr activation 55, which 1618 is itself correlated with late exponential/early stationary phase growth 54, the rapid 1619 depletion of AI-2 from local extracellular space leading to cell redispersion could be seen 1620 as part and parcel of a larger coordinated behavior of a population seeking out new 1621 energy sources to exploit. Alternatively, as Lsr signaling appears to mediate both the 1622 influence of AI-2 on biofilm development 60 and its consumption as a secondary 1623 metabolite 46, partial population activation in this context may represent a mechanism for 1624 the alignment of secondary metabolism consumption with the metabolic burden of 1625 producing public goods. As Lsr activation loosely corresponds with growth phase 1626 transitions, one specific instance of this concept is the possibility that chemotaxis to self-1627 secreted AI-2 draws a population toward surfaces at which preparations for a sessile 1628 93 lifestyle are initiated by the QS activated population fraction while the remaining cells 1629 continue to seek out other fertile grounds for colonization. 1630 Aside from phenotype diversification within a population, the ability to move 1631 toward a favorable growth environment and absorb signaling molecules from the 1632 extracellular space could have multiple motivations. For example, influx of AI-2 via Lsr 1633 activation could be used to limit QS sensing processes in competitors — processes 1634 involved in the transition to quiescence or sessility including: increased antibiotic 1635 resistance in local bacteria 138; increased predation defense 139,140, by preventing other AI-2 1636 consumers from competing for a similar niche; exploiting available secondary 1637 metabolites 46, or some combination thereof. Additionally, as an established population, 1638 competition for a secondary metabolite might be an expensive proposition — especially 1639 if the metabolite or its breakdown products moonlight as epistatic signals for the 1640 transition to stationary phase phenotypes. However, denying potential competitor 1641 populations secondary metabolite and/or limiting the prevalence of potential chemotactic 1642 cues could be a worthwhile tradeoff, particularly if only a select fraction of the 1643 population is required for the job. 1644 Whichever if any of these possibilities obtain for any particular circumstance, 1645 they all fall under the broad umbrella of hoarding amid the relative phylogenetic 1646 abundance of AI-2 in the microbial world 27. To more carefully consider how changes to 1647 the Lsr system itself might inform hoarding dynamics in consortia, mixed population 1648 simulations were performed. 1649 94 In mixed simulations, we found that luxS mutants remained largely inactive when 1650 paired with wildtype cells. This is attributable to a zero baseline of intracellular AI-2 in 1651 such cells, thus requiring a much higher extracellular threshold of AI-2 prior to 1652 activation. These populations represent a type of QS cheat termed “signal-negative” 141. 1653 Such cells are metabolically unburdened by the production of autoinducer even though 1654 they participate in the production of social goods and enjoy their benefits. In this 1655 particular case, the luxS population was also unable to compete for AI-2 against signal-1656 producing wildtype cells, and remained largely QS inactive. Thus, in mixed cultures, 1657 luxS negative cells were effectively signal blind as well. As long as QS products are 1658 public goods, luxS mutants are likely to be doubly non-cooperative. If the benefit of 1659 activation is direct or the resulting product is not a public good for the consortia, such a 1660 defect would likely minimize signal-negative cheating. This incentive for cells to avoid 1661 signal-negative cheating is compounded by any metabolic burdens arising from an 1662 incomplete methyl cycle 46, creating an incentive for cooperation even more direct than 1663 those previously considered for generic QS processes 142. Alternatively, luxS negative-1664 Lsr positive cells could play a niche role within bacterial consortium, consuming AI-2 1665 made by competing species 89. 1666 LsrF and lsrG double mutants were chosen to study in part because they are the 1667 most commonly missing elements among Lsr homologs 14. They were also found to be 1668 more sensitive to AI-2 in synthetic biology contexts 131, internalizing and accumulating 1669 AI-2 more rapidly than wildtype cells. lsrFG cells thereby suppressed wildtype 1670 activation. In general, lsrFG mutants present a sort of paradox. Even though these 1671 populations may be slightly advantaged by a faster activation and corresponding AI-2 1672 95 internalization, they are unable to utilize this increased pool metabolically, resulting 1673 merely in a higher AI2-P intracellular steady state. This may be warranted if the only 1674 role of AI2-P is to effect multiple downstream phenotypes via pleiotropic LsrR 1675 derepression. However, the degree to which LsrR derepression upregulates non-Lsr 1676 genes remains unknown 143. 1677 Alternatively, as representatives of other strains or species, interactions of lower 1678 AI-2 sensitivity strains (luxS) and higher AI-2 sensitivity lsrFG) strains with wildtype 1679 cells suggest that consortial Lsr activation for cell types with different Lsr/AI-2 sensitivity 1680 groups is likely to be highest for any particular subset when all constituent cells share a 1681 similar or lower Lsr/AI-2 sensitivity. This adds another layer of complexity to the 1682 efficiency sensing view of QS, where bacteria use QS machineries to evaluate whether 1683 dependent secreted public goods will serve their intended purpose, insofar as Lsr QS may 1684 allow interrogation of the environment for other cells with varied Lsr/LuxS affinity 1685 types. This potential outcome is one of many stemming from the intercellular negative 1686 feedback of the Lsr system operating even amid simple consortia as studied here. 1687 1688 1689 1690 1691 1692 96 3.6 Concluding Remarks 1693 While the simulations contained herein attempt to model quorum sensing 1694 contextually, the extent to which our in silico experiments correctly predict emergent 1695 behaviors must be considered 125. 1696 As a general matter, the reported results rely upon the validity of assumptions 1697 regarding the sensitivity of QS and chemotactic behaviors to AI-2. As a rule we have 1698 tried to make assumptions that are supported by the literature. For example, the minimal 1699 concentration at which AI-2 triggers chemotactic behaviors appears to be lower than that 1700 at which QS can be triggered for wildtype Lsr systems 131. Even if both behaviors were 1701 triggered at similar concentrations, clustering phenomena still likely obtain due to time 1702 lags associated with transcription, translation, and rate-limited transport through low flux 1703 pathways. 1704 Additionally, as previously mentioned, QS phenomenon apply only where 1705 extracellular transport is dominated by diffusive processes, when convection is minimal; 1706 otherwise, autoinducer dilution prevents activation 144. Therefore, QS has often been 1707 reframed as diffusion or efficiency sensing 145. Regardless of these semantics and 1708 connected interpretations, these phenomenon are likely to prevail within a protected 1709 volume or cavity with restricted access to flow conditions. 1710 Of further complication is the fact that the greater the diffusivity of the AI-2, the less 1711 pronounced chemotactic phenotypes are likely to be, as gradients rapidly dissipate. Also, 1712 the greater the diffusivity of AHL, the slower the colony will be to activate. While 1713 diffusivity coefficient values were specific to AI-2 and AHL, they were estimates based 1714 97 on Wilke-Chang correlation calculations146. However, while increasing diffusivity would 1715 blunt chemotactic behaviors, bimodality would be exacerbated as the effective distance 1716 between cells became smaller. An additional concern regarding the extracellular 1717 transport of chemical species is that bacterial movement might be reasonably expected to 1718 result in superdiffusive conditions directly proximal to the cell membrane. Due to the 1719 low Reynold’s number at these scales, however, this superdiffusivity should not extend 1720 into the surrounding bulk. 1721 Moreover, when asserted noise is tuned lower, Lsr bifurcation is also likely to 1722 dissipate. While the heterogeneity chosen is probably conservative, studies of population 1723 heterogeneity that track a significant number of single cell expression histories were 1724 unable to be identified. 1725 Finally, from a purely qualitative perspective, swimming populations generally 1726 appear to slow upon entering late-exponential or early-stationary phase growth. That is 1727 to say that the average speed of propulsion becomes markedly distributed, in a manner 1728 probably associated with growth phase heterogeneity. Commonly, in such 1729 circumstances, macroscopic self-propulsion comes to a halt altogether for some cells. As 1730 QS activity is often associated with growth phase transitions, it is beguiling to imagine a 1731 scenario where chemotaxis draws a population toward surfaces at which sessile behaviors 1732 are initiated by the less mobile population upon QS activation, while cells that still can 1733 respond chemotactically are less likely to be QS activated and continue to seek out more 1734 fertile grounds for colonization. 1735 1736 98 3.7 Supplemental Information 1737 3.7.1 Text and Discussion 1738 The focus of the main text, the Lsr system, is one of two known signaling transduction 1739 systems that responds to the ‘universal’ quorum sensing signal, autoinducer-2 (AI-2). In 1740 E. coli 60, Salmonella 77, and A. acetinomycetemcomitans 67, Lsr based QS drives or at 1741 least influences cooperative behaviors among multicellular consortia of bacteria. 1742 Experiments from Tsao, et al. 94 used a two plasmid system to amplify Lsr based 1743 expression by transcribing T7RNAP under pLsrR control on a single copy plasmid and 1744 linking GFP expression to a T7 promoter on a mid-to-high-copy pET200 plasmid. By 1745 one interpretation, this created a reporter system for Lsr QS activation 144,147. 1746 Interestingly, pure cultures of reporter transfected-wildtype bacteria developed a 1747 fluorescence distribution similar to that of mixed cultures. This suggested that one 1748 fraction of the wildtype population became QS active, while the other remained off in a 1749 bimodal fashion, belying the general association of QS with whole population 1750 coordination. 1751 Previous studies have modeled the prototypical LuxIR QS system as producing a 1752 strong switching behavior given sufficiently high cell density 127, usually attributed to the 1753 positive feedback inherent in the system topology 148. The Lsr system too, has been 1754 shown to generate behavior reflecting a highly sensitive switch 112. Modeling of Lsr 1755 activity in distinct cells within a population has not previously demonstrated lasting 1756 bimodality of system expression 95, however. 1757 99 Unlike the LuxIR system, Lsr topology resembles that of sugar importer systems 1758 like the lac operon, which has been shown to produce bimodal activity when exposed to a 1759 nonmetabolizable inducer 149. While thiomethyl galactoside and isopropyl--D-1760 thiogalactopyranoside served as a constant stimulus in those experiments, sufficient 1761 production of AI-2 throughout the time course of interest could serve as an analog in the 1762 case of the Lsr system. 1763 Using parallel ODEs to represent individual cells drawing from the same 1764 extracellular AI-2 pool, our model suggested that one subpopulation could prevent 1765 another from activating when paramet+ers were sufficiently perturbed. This analysis was 1766 extended to demonstrate that drawing parameter values from log normal distributions 1767 with different variances resulted in a range of partially activated populations. 1768 Additionally, our model suggested that noise from spatially associated stochasticity was 1769 by itself insufficient to generate bimodal expression patterns. 1770 3.7.2 Methods 1771 3.7.2.1 Lsr ODE Model 1772 3.7.2.1.1 Generalities and scope. The model described herein was developed to simulate 1773 Lsr system behavior for bacteria in a batch reactor between lag and early stationary 1774 growth phases. At the endpoint, AI-2 should be functionally depleted from the 1775 supernatant 54,55,57, with extracellular concentrations peaking between 4-6 hours 150. The 1776 developed equations relied on Hill and Michaelis-Menten like expressions to encapsulate 1777 reaction rate behaviors. In order to investigate population bifurcation, separate sets of 1778 100 ODEs were run simultaneously, modeling multiple cells sharing the same extracellular 1779 space. 1780 Lsr system components were modeled as translated from polycistronic mRNA 1781 species lsrRK and lsrACDBFG, the transcription of which was considered a function of 1782 LsrR and intracellular AI-2 concentrations. The proteins LsrK and LsrR were modeled as 1783 distinct species, LsrF and LsrG were modeled as a single entity. LsrA, LsrC, LsrD, and 1784 LsrB were folded into a single type of entity, OP, as they form a complex ABC type 1785 importer. Additionally, “LsrFG”, for while their products are distinct they both use AI2-1786 P as their substrate and neither’s products are known to feed back onto Lsr activity. Thus 1787 LsrF and LsrG were functionally equivalent for our specific purposes 54. Of the proteins 1788 outside the divergent Lsr operons that influence Lsr activity in E. coli (luxS, ydgG, crp, 1789 and pts), none are themselves modulated by Lsr activity 55,58,63. Moreover, the 1790 concentration of these species have not been shown to markedly vary within the time 1791 frame of interest. These proteins were therefore considered constant and their activity 1792 rates were simplified to a Michaelis-Menten like dependence on the substrate alone. 1793 3.7.2.1.2 mRNA expression. Production of mRNA species was modeled with a modified 1794 Hill equation derived from an expression for the fraction of free DNA to total DNA (free 1795 DNA + repressed DNA) as a function of LsrR and AI2P concentrations. Whereas active 1796 LsrR was modeled as a tetramer 151, AI2P cooperativity was asserted. Cooperativity for 1797 AI2P derepression was assumed to be on the order of cooperativity for protein/DNA 1798 binding. A lower bound for such cooperativity might be 1.38-2.72, which was measured 1799 for a nonspecific interaction between DNA and a multidomain protein in Mycobacterium 1800 smegmatis 152. An upper bound for such interactions is believed to be around 10 153. As 1801 101 each LsrR has a distinct AI2P binding domain, we chose 4 as a baseline level of 1802 cooperativity. The trajectory of the system using this cooperativity was similar to that for 1803 degrees of cooperativity greater than 4. At tested cooperativities less than 4, the drop in 1804 activation sharpness was sufficient to noticeably dull the rate of Ai-2 1805 recompartmentalization from the extracellular space. The transcription rates for the two 1806 different polycistronic mRNA species were set as equivalent, as the data on the relative 1807 strength of expression toward LsrA and LsrR is inconsistent 56,70. 1808 3.7.2.1.3 Protein synthesis. Lsr proteins expressed from the same polycistronic mRNA 1809 species were modeled as having the same translation rate. While different ribosome 1810 binding sites along the polycistrons are likely to have different affinities for ribosomes, in 1811 the absence of data, the co-regulation implied by operon structure and protein function 1812 was given overriding consideration. 1813 3.7.2.1.4 mRNA and protein degradation. As an additional simplification, both 1814 polycistronic mRNA species were modeled with the same rate of degradation, with a 1815 half-life on the order of 10 minutes, which was an order of magnitude faster than that 1816 used for proteins. While Lsr proteins are expected to have different vulnerabilities to 1817 proteolytic degradation, the result is either higher or lower quasi-steady state levels. The 1818 effects of such differences could be accounted for by either lowering or increasing 1819 corresponding unconstrained enzymatic rates. 1820 3.7.2.1.5 Cell growth. Cell growth was modeled as Monod growth. The maximum rate 1821 and Monod saturation rate constants of 0.032/min and 75 M, respectively, were had 1822 from a fit to OD600 measurements from a batch growth at 37oC. The growth rate also 1823 102 informed the dilution of cellular components. As a generality, the rate of cell growth 1824 strongly influenced simulation results, as increasing cell density accelerated extracellular 1825 AI-2 accumulation. That is, faster growth decreased the time to autoinduction. 1826 3.7.2.1.6 AI-2 transport. Instantaneous concentration and dilution associated with 1827 transport between extracellular and intracellular spaces was treated by including a 1828 dilution/concentration term of 1012 as seen in equation (10). This dilution factor 1829 accounted for the difference between the femtoliter intracellular environment and the 1830 milliliter term associated with cell concentration. 1831 Influx into the periplasm through porins was modeled as a diffusion process with 1832 a Michaelis-Menten form, rate limited at high concentrations, but otherwise proportional 1833 to the concentration difference between extracellular and periplasmic compartments. 1834 Influx through Lsr ABC type complexes was modeled as a function of importer complex 1835 concentration and periplasmic AI-2. The transporter’s component proteins were treated 1836 as a single species. Transporter complex was assumed to form with a cooperativity of 4, 1837 as a reflection of the fact that four independent components are involved in complex 1838 formation, namely LsrA, LsrB, LsrC, and LsrD. Although the nucleotide binding 1839 component, LsrA, operates as a dimer, any additional degree of freedom was considered 1840 to be eliminated by the fact that the LsrA dimer pair is fused together. Further, this 1841 transporter complex was asserted to act upon periplasmic AI-2 according to a Michaelis-1842 Menten like dynamic. Michaelis-Menten like dynamics were also asserted for the 1843 operation of the alternative importer as well as for AI-2 export through YdgG. That is, 1844 these processes were assumed to be constrained by a maximal velocity above a saturating 1845 103 AI-2 concentration, and that at lower AI-2 concentrations activity was approximately 1846 linear with respect to AI-2 itself. 1847 3.7.2.1.7 AI-2 degradation and synthesis. The rate of extracellular AI-2 degradation was 1848 assumed to be minimal based on a lack of attenuated AI-2 activity in bioassays after 1849 incubation of in vitro synthesized samples overnight at 37oC. The rate of cytoplasmic 1850 AI-2 degradation was set significantly higher to account for experiments measuring 1851 cytoplasmic AI-2 in E. coli without functional Lsr activity, wherein cytoplasmic AI-2 1852 concentrations dropped significantly once the stationary phase had been well established 1853 150. While the time frame of this marked decrease fell outside the scope of in silico 1854 experiments herein, it nonetheless suggests the existence of mechanisms that degrade AI-1855 2 intracellularly, independent of Lsr system expression. Periplasmic degradation of AI-2 1856 was modeled as intermediate to the extracellular and cytosolic rates, and did not bear 1857 strongly on the activity of the Lsr system, since the absolute moles in periplasmic pools 1858 were limited compared to extracellular and intracellular species. 1859 The same experiments that provide evidence for this degradation also suggest that 1860 the rate of synthesis is not constant 150, even if the expression of luxS and pfs are not 1861 strongly varied across time 56. Nonetheless, we model AI-2 synthesis as constant, which 1862 over the course of the time scale of interest we assume to be operationally approximate. 1863 3.7.2.1.8 Equations 1864 The specific form of the equations used and the parameters values used herein 1865 were as follows: 1866 1867 104 1868 (1) max [ ] [ ] [ ] [ ] [ ] [ ] ( [ ] ) ( ( ))[ ] [ ] peri out peri porin porin out peri pericoop in AIperi peri ABC peri d AI AI AI V dt Km AI AI AI V OP basal t AI Km AI            , 1869 (2) [ ][ ] ( [ ] ) [ ] [ ] [ ] [ ] ( ( ))[ ] [ ] [ ] pericoopin synth in ABC peri in in Phos ydgG AIin in Phos in export in AId AI K V OP basal dt Km AI AI AI k LsrK V t AI Km AI K AI             , 1870 (3) 2 [ ][ 2 ] [ ] [ ] [ 2 ] [ ] ( ( ))[ 2 ] [ 2 ] in Phos Phos in Cat AI P Cat AId AI P k LsrK dt Km AI AI P k LsrFG t AI P Km AI P         , 1871 (4) [ ] [ 2] ( ( ))[ ]P Td OP k mRNA t OPdt     , 1872 (5) [ ] [ 2] ( ( ))[ ]P FGd LsrFG k mRNA t LsrFGdt     , 1873 (6) [ ] [ 1] ( ( ))[ ]B Rd LsrR k mRNA t LsrRdt     , 1874 (7) [ ] [ 1] ( ( ))[ ]B Kd LsrK k mRNA t LsrKdt     , 1875 (8) 1 12 2 3 2 [ 1] 1 ( ( ))[ 1] [ ]1 [ 2 ] mR MRcoop coop d mRNA V t mRNA k LsrRdt r AI P       , 1876 (9) 2 22 2 3 2 [ 2] 1 ( ( ))[ 2] [ ]1 [ 2 ] mR MRcoop coop d mRNA V t mRNA k LsrRdt r AI P       , 1877 (10) [ ] [ ][ ] [ ]( [ ] [ ] [ ] ) [ ] peri outout max porin porin out peri in ydgG export in AI AId AI dilutionF cells V dt Km AI AI AI V K AI       , 1878 (11) [ ] [ ][ ] [ ]mid d cells yield Substrcellsdt K Substr  , 1879 (12) [ ] [ ][ ] [ ]mid d Substr yield Substrtake cellsdt K Substr   1880 105 , where [Substr] represents the concentration of substrate, [AIperi] represents the 1881 concentration of periplasmic AI-2, [AIin] represents the concentration of cytoplasmic AI-1882 2, [AIout] represents the concentration of extracellular AI-2, [AI2P] represents 1883 phosphorylated AI-2, [LsrR] represents LsrR concentration, [LsrK] represents LsrK 1884 concentration, [LsrFG] represents LsrF and LsrG concentration, [mRNA2] represents 1885 lsrRK concentration, [mRNA1] represents lsrACDBFG concentration, and [OP] 1886 represents transporter protein concentration. 1887 3.7.2.1.9 Parameter Values. A table of parameter values can be found in Table 3-S1. 1888 The maximal rate of transport through the alternative importer, as governed by basal, was 1889 set lower than the rate of export through YdgG, allowing extracellular AI-2 accumulation 1890 at baseline levels of Lsr activity. The remaining parameters were set such that this initial 1891 AI-2 flux out was overcome by Lsr activation. 1892 Specifically, according to BB170 bioassays of culture supernatant, AI-2 appears 1893 to peak between 4-5 hours in a batch culture. FRET-LuxP assays and bioassays suggest 1894 that the peak concentration is less than 80 M and greater than 40 M, respectively 55,150. 1895 Once extracellular accumulation stalls, AI-2 appears to be drawn down to low 1896 concentrations of AI-2 in less than an hour and then to concentrations below bioassay 1897 sensitivity within the next hour. 1898 Asserting that transcription is directly coupled to translation for LacZ, 1899 transcription was fit to Miller assays of LacZ expression from pLsrA and pLsrR 1900 promoters, indicating that transcription begins evolving prior to 4 hours after culture 1901 initiation, reaching a several fold higher level of expression prior to 6 hours 56. 1902 106 Table S1. Estimated parameter values 1903 Parameter Value Parameter Value Parameter Value 8000 M/min 0.15 /min 7x10 -5 M4 1 M 50 /min 4 2x10 12 /(min*M3) 1.4 M 6 4 18.8 /min 0.693 /min 500 M/min 0.01 /min 0.693 /min 0.5 M 0.01 /min 10 -12 0.015 /min 18.8 /min 0.3 /min 220 M/min 0.01 /min 0.032 /min 80 /min 0.01 /min 2250 M 1.4M 2x10 -5 M/min 4.5x10 -7 [Substr]/[cell] 1250 M/min 2x10 -5 M/min 0.5 M 3.5x10 8 1904 1905 1906 1907 1908 1909 1910 1911 107 While the selected parameter set in combination with the given ODE’s produced a 1912 much faster depletion of extracellular AI-2 than was observed experimentally, the 1913 fraction of QS induced cells evolved over an extended period of time 94, and it was this 1914 evolving population’s averages to which the model was fit (Figure 3-S1). In order to 1915 generate an evolving fractional Lsr autoinduction, a tractable number of cells was 1916 modeled within a finite difference agent based scheme. Each cell’s processes were 1917 modeled by its own set of ODEs and the cell population was slightly desynchronized in 1918 order to generate the fractional induction seen by Tsao, et al 94. 1919 A parameter search was carried out to identify parameter sets satisfying the above 1920 criteria. Among other measures of fit, the time to Lsr autoinduction was sensitive to 1921 changes in all parameters effecting transport due to their direct bearing upon the balance 1922 of accumulating cytoplasmic AI-2 and corresponding AI2-P species. The parameter 1923 space satisfying the available data including extracellular AI-2 concentration, was 1924 nonetheless broad with similar sensitivities and behaviors over a range of values. 1925 3.7.2.1.10 Initial Values. Initial values were set according to approximate steady state 1926 values for a system without lacking AI-2 production, where the Lsr system was 1927 uninduced: [cells] = 3e7, [Substr] = 1500 M, [AIperi] = 0.015 M, [AIin] = 0.2 M, 1928 [AIout] = 0.02 M, [AI2P] = 0.03 M, [LsrR] = 0.0012 M, [LsrK] = 0.0012 M, [LsrFG] 1929 = 0.0012 M, [OP] = 0.0012 M, and [mRNA2] = [mRNA1] = 0. 1930 1931 108 3.7.2.1.11 Numerical solutions. The equations and parameters were solved using 1932 NDSolve in Mathematica 8.0 utilizing “StiffnessSwitching” methods. The solution 1933 required the “StiffnessSwitching” option, without which, the solution was unstable. 1934 3.7.2.2 Finite difference-agent based model 1935 3.7.2.2.1 Modeled Environment. The environment was defined as a 500 x 500 x 6 m 1936 volume, and was divided into 2 x 2 x 6 m elements. Cells either exported or imported 1937 AI-2 from their intracellular space into or from the finite difference element in which 1938 their cell centers were found. AI-2 also diffused between finite difference elements as 1939 modeled by a forward in time-central in space scheme with an assumed diffusion 1940 coefficient of 5x10-7 cm2/s. The boundaries of the simulation were modeled as 1941 impermeable. 1942 3.7.2.2.2 Adaptation of equations and solutions. With few exceptions, the previously 1943 described ODEs were repurposed without modification in the finite difference/agent 1944 based model. Growth was one exception, as division became a discontinuous stochastic 1945 event. Monod growth dynamics informed the median of a log normal distribution of 1946 doubling time with a  of 0.05 where the median rate was updated every time step. The 1947 exchange of AI-2 between the environment and a cell was localized to the space in which 1948 the cell center was found at the beginning of the time step. This allowed a synchronized 1949 update of the final grid concentration at the end of the time step. Furthermore, the AI-2 1950 dilution/concentration factor was adjusted from 1012 to 24 to account for the difference 1951 between the milliliter volume associated with cell concentration and the implied volume 1952 of grid elements. 1953 109 3.7.2.2.3 Cell ODE Numerical Solution Method. The numerical method used in the 1954 agent based modeling was a second order Runga-Kutta with an ad hoc allowance for 1955 stiffness. In order to achieve efficient calculation, we used an explicit method with the 1956 exception of periplasmic AI-2 after transporter concentration exceeded 4 times its initial 1957 concentration. This threshold was chosen based upon NDSolve interpolation, as a point 1958 at which the Lsr system including transporter expression and subsequent concentration 1959 had already begun its transition to an active state. After this threshold was surpassed, 1960 porins were assumed to be the rate limiting element in the transport of AI-2 from the 1961 extracellular space to the cytosol, and modeling was made to reflect this by substituting 1962 terms describing ABC-type transporter activity with terms describing porin activity; 1963 furthermore, periplasmic AI-2 was held at zero, reflecting the numerical solution from the 1964 ODE system described in previously. 1965 3.7.2.2.4 Cell Division. Cell division was governed by individual counters that 1966 incremented at each time step. Once a cell’s counter exceeded its doubling time, division 1967 occurred. At the time zero, cell counters were set randomly between zero and the 1968 maximum allowed value. Doubling time was varied between cells according to a mean 1969 growth rate based on Monod kinetics (with parameters found in Table S1) with a log 1970 normal distribution (variance of 0.05) in order to desynchronize cell doubling. Upon 1971 division, both mother and daughter cells acquired new growth rates from a log normal 1972 distribution, while bearing duplicate properties including initial position, and their age 1973 counters were reset to zero. Newly initialized cells were assigned a basal rate of AI-2 1974 flux through alternative importer pathways (the parameter, basal) or a rate of AHL 1975 synthesis, for Lsr or LuxIR simulations respectively. 1976 110 3.7.2.2.5 AI-2 Diffusion. AI-2 diffusion was modeled using a central difference 1977 approximation. While both truncation and roundoff error arise from this process, the 1978 overall behavior of the simulation was not expected to be dramatically impacted, as 1979 clustering behavior also obtained when elements were four times the size. The diffusion 1980 coefficient used (~5x10-7 cm2/s) was approximated using the Wilke-Chang correlation 146. 1981 3.7.2.2.6 Time interval ordering. The order of calculations within a time step was as 1982 follows. First, the average growth rate was determined from a Monod growth dynamic 1983 solved by a second order Runga-Kutta method as a function of substrate concentration 1984 and E. coli density. Second, each E. coli divided or did not divide, marginally 1985 accommodated its AI-2 chemotactic threshold (if appropriate), and moved (according to 1986 the particular scheme employed). For the specific purposes of the simulations reported 1987 here, cells moved randomly in space each time step at an average rate of 20 m/sec, 1988 coming to an average distance of 1.3 m per time step. Correspondingly, cells rarely 1989 moved more than one grid element at a time for any given time step. Finally, each grid 1990 was subject to calculations approximating diffusion. That is, AI-2 was allowed to move 1991 according to molecular diffusivity and distance/time calculations. After all diffusion 1992 processes were calculated over the entire environment, the exchange of AI-2 between 1993 bacteria and the grids in which the bacteria were present was calculated according to Lsr 1994 dynamics governed by previously described ODEs. The sum of these changes to AI-2 1995 concentration were then applied to each grid and each cell. Time step and grid size were 1996 chosen such that moderate changes to these measures resulted in qualitatively indistinct 1997 outcomes. 1998 111 All random assignments were drawn according to the Mersenne Twister algorithm 1999 154. Seeds for the algorithm were changed every simulation. 2000 3.7.3 Results 2001 3.7.3.1 Numerical solutions to ODEs. Given the ODEs and parameter values used 2002 herein, the system evolved from a state of low expression to a state of high expression 2003 (QS activated), as can be seen in Figure 3-S1, where the trajectories of different state 2004 variables are represented, normalized to their maximum values. This corresponds to the 2005 known pattern of Lsr activity development. At a particular AI2-P-to-LsrR threshold, 2006 system expression is de-repressed, leading to the increased expression of transporter 2007 proteins, which itself leads to an influx of extracellular AI-2 and a sustained increase in 2008 the expression of lsr mRNA species. 2009 In the model, heightened activity appeared to persist even after the depletion of 2010 extracellular AI-2, by virtue of an altered flux balance from net AI-2 exportation to one 2011 of net importation of processing. That Lsr autoinduction leads to a shifts in the 2012 compartmentalization of AI-2 species from one dominated by extracellular species to one 2013 dominated by cytosolic species is one interpretation of Figure 3-S2. 2014 3.7.3.2 System Sensitivity to basal rate of AI-2 uptake. As expected, system behavior 2015 was sensitive to parameters that markedly affected the rate of AI2-P accumulation when 2016 Lsr system expression was low. Among these was the parameter basal (Figure 3-S3). 2017 With increasing value, the rate of AI2-P accretion increased, accelerating the time 2018 to Lsr activation. 2019 112 2020 Figure 3-S1. Comparison of solution for population with a single basal value versus 2021 a population with a unimodally distributed value of basal. Juxtaposition of the 2022 solution for extracellular AI-2 for a simulation of cells with a single basal value versus 2023 the average solution of extracellular AI-2 for a simulation of cells with a log normal 2024 distribution of the parameter basal 2025 2026 2027 2028 2029 2030 2031 2032 113 2033 Figure 3-S2. Numerical solution to developed ODE equations. A Selected 2034 interpolated trajectories of the solution to the ODEs from 3.3.1.7 with the parameters 2035 from 3.3.1.8. B Mole fraction of total AI-2 species located either extracellularly or 2036 cytosolically. The denominator was the mole sum of extracellular, periplasmic, and 2037 cytosolic AI-2. 2038 114 2039 Figure 3-S3. Parameter sensitivity to the parameter basal as seen by changes in the 2040 time until system activation. System trajectories given different values of the parameter 2041 basal. 2042 2043 2044 2045 2046 2047 2048 115 Specifically, an ~8% increase in the rate of AI-2 influx through the alternative pathway 2049 from the base value resulted in a ~21% reduction in the time to activation. The values 2050 shown in this figure represent a range of basal that corresponds roughly to that of ribose 2051 uptake via its low affinity mechanism. Or, correspondingly, that basal level denoted 530 2052 corresponds to 10% of the maximum achieved when the uptake mechanism is switched 2053 on. All these values (490-530) represent metabolically realistic values of small molecule 2054 uptake. 2055 3.7.3.3 Two sets of Lsr with different rates of basal AI-2 uptake. More importantly, 2056 when two sets of ODEs drew from the same pool of extracellular AI-2, one system of 2057 ODEs could prevent the other from activating given sufficient parameter perturbation, as 2058 seen in Figure 3-S4. As the perturbation dwindled, the two ODEs converged to the same 2059 high expression state. That is, two sets of ODEs, representing two different populations, 2060 were arranged to use the same extracellular AI-2. Solutions to ODE’s in Figure 3-S4A 2061 shared the same parameter values except for that of basal, which reflects the rate of AI-2 2062 flux into the cell from the periplasmic space through the alternative pathway. The value 2063 of basal was set to 510 for one population and was adjusted continually lower for a 2064 second, gradually slower activating, population. When the difference between the two 2065 values of basal was sufficiently large, the faster activating population depleted 2066 extracellular AI-2 prior to the slower population achieving a sufficient intracellular 2067 threshold to trigger Lsr induction, thus preventing QS activation. 2068 116 2069 Figure 3-S4. Solution to dual ODE system where second population had varied 2070 values of parameters related to the transport of AI-2. Population 1’s parameter values 2071 were held constant. A Over separate simulations, population 2’s parameter, basal was 2072 varied at comparatively lower values. The parameter values are in parentheses. The 2073 trajectory of AI2-P accumulation for each of these cases is represented here. B Over 2074 separate simulations, population 2’s parameter, Vin was varied at comparatively lower 2075 values. The trajectory of AI2-P accumulation for each of these cases is represented here. 2076 The values in parentheses reflect Vin/1012. 2077 2078 117 While basal represented a specific activity, the phenomenon of a faster Lsr 2079 activating group preventing a slower group from inducing proved a more general 2080 property of the model, holding when other parameters to which Lsr induction was 2081 sensitive were varied instead. This can be seen in Figure 3-S4B, where the variation in 2082 the assigned value of the parameter Vin, representing the rate of AI-2 transport through 2083 the Lsr ABC-type transporter, was substituted for variation in the value of basal. A 2084 decreasing series of values for Vin , all smaller than that of population 1, were assigned to 2085 population 2. Again, once the difference in Vin between populations became sufficient, 2086 population 1 was able to prevent Lsr activation in population 2. 2087 3.7.3.4 Full population of cells with Lsr in a finite difference environment. While 2088 coupling two systems of ODEs to the same pool of extracellular AI-2 allowed us to 2089 interrogate possible Lsr signal bifurcation of two different populations, we extended this 2090 analysis to include a full population of cells assigned basal values from a log-normal 2091 distribution instead of an effective binomial one. This analysis was conducted using 2092 mixed finite difference-agent based simulations, as described in the methods. 2093 As discussed in the main text, Lsr induction was bimodal, with the exact balance 2094 dependent on the variance of the basal distribution. This likely arose due to cells with 2095 higher values of basal from the leading edge of the distribution activating earlier and 2096 earlier as the variance of the distribution increased. The earlier the activation of these 2097 cells relative to the mean, the faster extracellular AI-2 species were drawn down, thus 2098 preventing the activation of cells with lower basal values, which themselves on average 2099 were liable to activate later and later. 2100 118 2101 Fig. S5 - Fraction of cell population QS activated decreases as the variation of the 2102 parameters Ksynth (A) and VydgG (B) increases. The influence of standard deviation of 2103 the natural logarithm, , of the log-normal distributions on the fraction of the population 2104 that was QS activated is represented here for standard deviation values ranging from 0 to 2105 0.025, each run in triplicate. Standard deviation of the natural logarithm for the log-2106 normal distribution was used because the coefficient of variance is solely dependent on 2107 this measure. Dark lines represent the average values, whereas lighter surrounding lines 2108 represent the standard deviation. 2109 119 Apart from basal, separately varying Ksynth and VydgG , the rate of AI-2 synthesis and AI-2 2110 export, respectively, achieved similar changes to the bimodal nature of Lsr activation 2111 (Figure 3-S5). Although activation levels plateaued at different fractions of the 2112 population when the same standard deviations were applied to the distribution of these 2113 parameters, increased variation in both Ksynth and VydgG consistently decreased the fraction 2114 of the population that was ultimately activated. This serves to further indicate that 2115 population level bimodal expression of the Lsr system may be a function of any 2116 heterogeneity that desynchronizes AI-2 recompartmentalization. 2117 3.7.3.5 Minimal role of spatial heterogeneity of AI-2. In addition to evaluating the 2118 role of heterogeneous expression at the population scale, whether spatially associated 2119 stochasticity might influence bimodal expression was also inferred, mainly from 2120 comparison of simulations using a standard finite difference scheme against simulations 2121 where the entire environment was defined by a single element. Treating the environment 2122 homogenous made all AI-2 simultaneously available to all cells, whereas cells only 2123 interacted with AI-2 in their own element using a standard finite difference approach. As 2124 shown in Figure 3-S6, the approaches yielded highly similar AI2-P trajectories when cell 2125 motility was undirected. This was the case for all state variables modeled. Furthermore, 2126 in standard finite element environments, when governed by a single parameter set, cell 2127 populations became wholly activated over a very small window (Figure 3-S5,  = 0). If 2128 heterogeneity arising from spatial stochasticity influenced the bimodal phenotype, 2129 population activation would be expected to be incomplete. The absence of such an effect 2130 implied that spatial stochasticity did not play a marked role in shaping bimodal response. 2131 120 2132 Figure 3-S6. Comparison of results from single versus multiple finite difference 2133 elements to define environment. The average trajectory of AI2-P for cells with the 2134 same parameter sets in simulations where the environment was defined as either a single 2135 finite difference element or by the standard array of elements as defined in the methods. 2136 Modeling with a single finite difference element eliminates spatial noise as a source of 2137 difference between cells. The addition of noise through the full implementation of finite 2138 difference elements, adds spatially associated noise to the simulation. This did not result 2139 in a significant change in the average trajectory of AI2-P. 2140 2141 2142 2143 2144 2145 121 3.7.3.6 Agreement between numerical ODE and finite difference agent based 2146 solutions. As a general comment, the agreement between numerical ODE solutions and 2147 the finite difference-agent based approach was inexact. In particular, the time to 2148 activation was offset between the two solutions as seen in Figure 3-S7A. Nonetheless, 2149 the solution trajectories were similar and an evaluation of the time to activation as a 2150 function of basal indicated that parameter sensitivities between the solution approaches 2151 were congruous, as seen in Figure 3-S7B. 2152 3.7.3.7 Heterogeneity of local Lsr and LuxIR QS activation in growing colonies. 2153 Smoothing of heterogenous input by LuxIR QS was not shared by Lsr QS. This is 2154 implied in Figure 3-S8A, which shows the distribution of key parameters from each 2155 simulation type. For LuxIR simulations (left), the uninduced rate of AHL synthesis was 2156 varied among cells according to the depicted distribution. The higher the rate of synthesis 2157 for uninduced cells, the faster the QS activation. For Lsr simulations (right), the basal 2158 rate of AI-2 influx was varied. Here too, the higher the rate of basal AI-2 influx, the 2159 faster the QS activation. The red line in both distributions represents the average 2160 parameter value for the first cell to activate from twenty simulations. For both Lsr and 2161 LuxIR simulations, the first cells to activate were on average all from the higher end of 2162 the imposed heterogeneity. For LuxIR QS (left), however, the first cells to activate were 2163 not always those with the highest rate of basal AHL synthesis, as indicated by the wide 2164 standard deviation. This is clearer when compared to Lsr QS (right), where the first 2165 activators were exclusively found at the tip of distribution for basal values (reflected in 2166 the limited variance). Essentially, while LuxIR/AHL activation smoothed out 2167 heterogeneity associated with the rate of AHL production, Lsr/AI-2 dynamics were 2168 122 2169 Figure 3-S7. Congruence of solution from finite difference-agent based modeling 2170 versus implicit solution of pure ODE’s. A AI2-P trajectory from implicit numerical 2171 methods and the average AI2-P concentration from the finite difference-agent based 2172 approach. Here, cells from the finite-difference-agent based solution all held the same 2173 parameter values as that from the pure ODE solution. In the pure ODE approach, cells 2174 were modeled as a dependent variable. Ideally, the two solutions would bear identical 2175 traces. B The rate to activation was assessed by fitting the function, f(t), from 12-152 2176 minutes to a first order linear regression, g(t). The first time point at which f(t)-g(t)>2g(t) 2177 was considered the point of activation. The time to activation for each value of basal was 2178 calculated and the bearing on the solution by the modeling and numerical method used 2179 was evaluated by direct comparison along the primary axis and according to the ratio of 2180 activation times for the finite difference-agent based solution to the pure ODE solution on 2181 the secondary axis. 2182 123 2183 Figure 3-S8. Measures of the difference between LuxIR and Lsr activation in the 2184 context of colony growth. A Histograms of randomly generated parameter values, each 2185 with an event count of 10,000. The average of the parameter values associated with the 2186 first cell to QS activate in each simulation and its standard deviation from among 20 2187 simulations are overlaid on the histogram in red. In LuxIR simulations, while the 2188 selection is biased, a significant number of cells with less than the median parameter 2189 value indicating that LuxIR has a smoothing effect on heterogeneity. B The dark lines 2190 represent the average local heterogeneity of 20 simulations, while the lighter, surrounding 2191 shades represent the standard deviation of those values. Inset is local heterogeneity of 2192 QS activation, where active cells were placed randomly within the colony as a function of 2193 percent of cells that were QS active for that colony (n=100). 2194 2195 2196 2197 2198 2199 124 unable to smooth similar non-genetic heterogeneity, instead producing population 2200 desynchronization. This difference in response to non-genetic heterogeneity was 2201 presumably a reflection of topological differences between QS signaling modules. 2202 This distinction between LuxIR and Lsr QS responses was also reflected in the 2203 spatial heterogeneity of activation. Local heterogeneity of QS activity was measured by 2204 using a heuristic that scored highest when every cells’ neighbors, up or down, left or 2205 right, were of the opposite QS state. That is, a value of one indicated a perfect 2206 checkerboard pattern of alternating activation, whereas a score of zero indicated that all 2207 cells were of the same QS state. That score was then averaged over the entire population 2208 to arrive at the score reported in Figure 3-S8B. For perspective, inset is a graph of the 2209 same measure for colonies whose QS active cells were randomly distributed, with the 2210 ordinate axis reflecting the likelihood that any one cells was QS activated. As the 2211 measure was only concerned with changes between QS state, scoring was symmetric 2212 about 0.5. 2213 At time zero, all simulations began with a score of zero (QS unactivated). Lsr/AI-2214 2 simulations were run with a median basal of 487.8 and a coefficient of variance of 2215 0.052 ( = 0.0225). For such an Lsr/AI-2 population, the first non-zero local QS 2216 heterogeneity values emerged near 120 minutes. The average of twenty simulations 2217 (dark blue line) and the standard deviation (surrounding light blue band) are depicted. As 2218 more cells became activated, local QS heterogeneity increased as both outlying and inner 2219 cells were induced. Local QS heterogeneity reached a plateau near 0.34 as QS activation 2220 began to abate. Given that the local QS heterogeneity for a cell colony wherein the same 2221 fraction of QS active cells were completely randomly distributed was 0.4 (inset), the local 2222 125 QS heterogeneity for Lsr activation was high, while retaining some degree of non-2223 randomness. This stood in contrast to the LuxIR case. LuxIR activation began near the 2224 colony center at 225 minutes, and increased rapidly, reflecting the rapid expansion of 2225 activation. However, spatial heterogeneity scoring of LuxIR QS activity peaked near 0.1 2226 despite a ~50% QS activation rate at that point. Cell colonies where 50% of the cells 2227 were QS active but where those cells were randomly placed within the colony averaged a 2228 heterogeneity score near 0.5. Low spatial heterogeneity associated with LuxIR/AHL 2229 activation is attributed to two primary factors: QS activity originates from a single 2230 centralized location and QS “On”/”Off” distinctions are quickly obliterated as QS active 2231 cells turn on their QS inactive neighbors. 2232 3.7.3.8 Evaluation of clustering when Lsr QS is coupled to AI-2 chemoattraction. 2233 Cluster and dispersal patterns were observable from inspection of Figure 3-S9, where the 2234 distance between cells across simulated time is shown for different swimming modes. 2235 Here, fully functioning cells (green, Lsr + chemotaxis) were compared to non-2236 chemotaxing populations (blue, randomly moving) or populations lacking the ability to 2237 recompartmentalize AI-2 (purple, lsr operon negative and non-specific uptake minus; Vin 2238 = 0, basal = 0). These alternative populations represented groups of cells that did not 2239 cluster or clustered but did not disperse, respectively. Initially, the median minimal 2240 distances for all population types were identical. The median minimal distance between 2241 cells decreased for all populations as a function of growth, as expected. However, the 2242 median minimal distance between cells for the two AI-2 chemotaxing populations 2243 decreased more rapidly than their non-taxis counterpart due to clustering. Cells unable to 2244 recompartmentalize (or uptake) AI-2 had a higher net flux out (they synthesize by do not 2245 126 2246 Figure 3-S9. Clustering of cells with lsr activity and AI-2 chemoattraction as 2247 measured by cell-cell distance. The median minimal distance between cells for (i) a 2248 population with both Lsr activity and AI-2 chemoattraction, (ii) a population lacking Lsr 2249 recompartmentalization ability, and (iii) another population lacking AI-2 chemoattraction 2250 but with recompartmentalization. Each condition was simulated twenty times, the average 2251 median minimal distance between cells is represented by the darker line, while the 2252 standard deviation of those values is represented by the surrounding lighter regions. The 2253 darker green box indicates the time over which the switch in median minimum distance 2254 between cells from that of an AI-2 chemoattracted population to one of a population not 2255 chemoattracted to AI-2 occurred for the wildtype population, while the lighter green 2256 region indicates the time over which Lsr induction began for the wildtype population. 2257 The bump around 250 minutes is indicative of this change. Inset is a pictorial example of 2258 the minimum distance between cells. 2259 2260 2261 2262 127 take in AI-2) than wildtype cells, achieving a higher extracellular concentration of AI-2 2263 over a longer distance. This may account for the slower or looser clustering of such cells 2264 compared to the wildtype phenotype. For wildtype AI-2 chemotaxing cells (Lsr+), the 2265 median minimal distance between cells began increasing around 225 minutes a short 2266 while after the population began to QS activate around 180 minutes. The delay was 2267 likely a function of the time required to begin drawing down the AI-2 concentration in 2268 these clusters, while the increasing distance between cells reflected the dispersion 2269 phenomena, which is apparent by 300 minutes. We note that the dispersion phenomena 2270 here is underrepresented locally because the calculated value of the intercell distance is 2271 averaged over the entire population, including all clusters and dispersed cells. In the end, 2272 the distance between wildtype cells matched the distance between cells lacking AI-2 2273 chemoattraction, indicating that the clusters had fully dispersed. While this is an exciting 2274 outcome that could have broad ramifications, we know of no directly aligned 2275 observations. 2276 3.7.3.9 Motility mode feedback onto population activation as a function of cell-cell 2277 distance. In general, we found that among the populations simulated here, non-taxis 2278 swimming populations were the slowest to QS activate. We note, however, that these 2279 same cells ultimately achieved the largest proportion of stably QS activated cells. This 2280 inverse correlation persisted across motility types (Figure 3-S10A). Cells in simulations 2281 of growing colonies activated the fastest but also experienced the smallest fraction of 2282 stably activated cells, whereas chemotaxing populations experienced intermediate levels 2283 of both speed to activation and of the final proportion activated. In other words, as a 2284 generalization, the higher the cell density the earlier the activation. However, higher 2285 128 density also appeared to produce stronger negative extracellular feedback from QS 2286 activation as a smaller fraction of the population was ultimately activated. (Figure 3-2287 S10B) 2288 2289 2290 2291 2292 2293 2294 2295 2296 2297 2298 2299 2300 2301 129 2302 Figure 3-S10. Measures of the difference between different modes of motility when 2303 coupled with Lsr/AI-2 dynamics. A The fraction of the population that was QS 2304 activated over time, in simulations of different motility, with average values (n = 20) set 2305 in a darker thinner line and a lighter surrounding shade representing the standard 2306 deviation. B The median minimum distance between cells for populations influenced by 2307 different combinations of motility and AI-2 uptake. Dark lines are an average value (n = 2308 20), while the surrounding lighter shades reflect the corresponding standard deviation. 2309 For example, cells undergoing colony growth had a predefined, regular distance between 2310 them, thus a single value prevailed across the entire time course and variability was zero. 2311 2312 2313 2314 2315 2316 2317 2318 2319 2320 130 Chapter 4: Conclusions 2321 Living within consortia is likely the primary mode of existence for bacteria. 2322 Therein, they must frequently negotiate numerous interactions with other species. While 2323 AI-2 is likely to be only one molecule within a vast sea by which such interactions are 2324 mediated, due to the prevalence of LuxS 46 and YdgG 59 homologs, it is one that is likely 2325 to be found in a plethora of ecological contexts. This prevalence alone makes it a likely 2326 candidate for QS and QS-like operations, the actual downstream functions of which are 2327 expected to be manifold. In an ecological context, QS and QS-like operations are likely 2328 to mediate the most types of cell-cell interactions, consisting of cooperation among and 2329 between species, coercion of one species’s action by another, and as cues from one 2330 bacteria to another without the intent of cooperation. Representing coercion, signal blind 2331 mutants within an otherwise cooperating community cause signal responding bacteria to 2332 QS. When the QS response involves the production of public secreted goods, signal 2333 blind mutants benefit without the metabolic expenditure and are coercing their signal 2334 responding counterparts into, at the very least, premature QS. Alternatively, purely as an 2335 indicator of cell density, autoinducers could potentially serve as a cue. 2336 The widespread phylogenetic signal of the Lsr system as described in the second 2337 chapter suggests that the Lsr system plays a functional role in numerous bacteria. The 2338 exact nature of that role is possibly varied from species to species, especially considering 2339 that it is unlikely that the affinity of LsrR for the intergenic region coevolved in an exact 2340 manner. As the Lsr system is known to affect biofilm development, it is believed to 2341 influence cooperative behaviors at least to this extent 60. In the third chapter, 2342 mathematically modeling the form native to E. coli, we lend further credence to the idea 2343 131 that the Lsr system leads bacteria to serve a dual role, as both a cooperator and coercer as 2344 a result of bimodal expression arising from population heterogeneity. 2345 We also demonstrated through modeling, that bimodality results in multiple 2346 emergent phenomenon depending on the mode of motility it is paired with, which itself 2347 fed back onto the bimodal phenotype in silico. As multiple homologs of the Lsr system 2348 exist, it is highly likely that three categories of Lsr homolog exist relative to E. coli’s 2349 system: more sensitive, equally sensitive, and less sensitive to AI-2. In terms of speed to 2350 activation these might roughly translate to: faster, equal, and slower to activate 2351 respectively. Here, we interrogated how bacteria containing such homologous systems 2352 interact with a population possessing a base E. coli Lsr system, and suggest lsrFG 2353 mutants, double lsrFG luxS mutants, and luxS mutants as representatives of each 2354 respective category. We showed how these competing populations could either largely 2355 inhibit, act in concert with, or cause wholesale activation of the wildtype population, 2356 respectively. For a luxS mutant population, the cost of coercion is the loss of the activated 2357 methyl cycle, indicating that at least for this case there is a built in disadvantage to free 2358 ridership. 2359 Placing the Lsr system within the larger context of other QS architectures, our 2360 modeling strongly indicates that consequent to bimodal activation, the patterns of 2361 expression arising from the Lsr system are in stark contrast to those associated with 2362 LuxIR QS. Moreover, if placed in the same environment as the AI-2 activated TCRS, 2363 LuxPQ, our studies indicate that Lsr activation would likely curtail LuxPQ based 2364 signaling coordination, at least up to a point. Indeed, in isolation the Lsr system operates 2365 more closely to sugar importation systems than other QS systems. Clearly, the 2366 132 distinction between sugar systems and the Lsr system is the context of self-production 2367 and self-secretion. Based on previous examination of QS 117,118, we believe that this 2368 secretion is likely to result in greater coordination compared to a population with the 2369 same average rate of AI2-P accretion but without YdgG or other means of AI-2 export. 2370 Aside from this consideration, our homology search indicates that the Lsr system 2371 phylogenetic signal is much less monophyletic than that for the lac system, even if it is 2372 more widespread. 2373 2374 2375 2376 2377 2378 2379 2380 2381 2382 2383 2384 2385 133 References 2386 2387 1. Greenberg, E., Hastings, J. & Ulitzur, S. Induction of luciferase synthesis in Beneckea 2388 harveyi by other marine bacteria. Arch. Microbiol. 120, 87–91 (1979). 2389 2. Eberhard, A. Inhibition and activation of bacterial luciferase synthesis. J. Bacteriol. 109, 2390 1101–1105 (1972). 2391 3. Fuqua, W. C., Winans, S. C. & Greenberg, E. P. Quorum sensing in bacteria: the LuxR-2392 LuxI family of cell density-responsive transcriptional regulators. J. Bacteriol. 176, 269 2393 (1994). 2394 4. Hagen, S. J., Son, M., Weiss, J. T. & Young, J. H. Bacterium in a box: sensing of quorum 2395 and environment by the LuxI/LuxR gene regulatory circuit. J. Biol. Phys. 36, 317–327 2396 (2010). 2397 5. Carnes, E. C. et al. Confinement-induced quorum sensing of individual Staphylococcus 2398 aureus bacteria. Nat. Chem. Biol. 6, 41–45 (2009). 2399 6. Boedicker, J. Q., Vincent, M. E. & Ismagilov, R. F. Microfluidic Confinement of Single 2400 Cells of Bacteria in Small Volumes Initiates High‐Density Behavior of Quorum Sensing 2401 and Growth and Reveals Its Variability. Angew. Chem. Int. Ed. 48, 5908–5911 (2009). 2402 7. Alberghini, S. et al. Consequences of relative cellular positioning on quorum sensing and 2403 bacterial cell‐to‐cell communication. FEMS Microbiol. Lett. 292, 149–161 (2009). 2404 8. Hense, B. A. et al. Does efficiency sensing unify diffusion and quorum sensing? Nat. Rev. 2405 Microbiol. 5, 230–239 (2007). 2406 9. Redfield, R. J. Is quorum sensing a side effect of diffusion sensing? Trends Microbiol. 10, 2407 365–370 (2002). 2408 10. Diggle, S. P., Gardner, A., West, S. A. & Griffin, A. S. Evolutionary theory of bacterial 2409 quorum sensing: when is a signal not a signal? Philos. Trans. R. Soc. B Biol. Sci. 362, 2410 1241–1249 (2007). 2411 11. Mascher, T., Helmann, J. D. & Unden, G. Stimulus perception in bacterial signal-2412 transducing histidine kinases. Microbiol. Mol. Biol. Rev. 70, 910–938 (2006). 2413 12. Case, R. J., Labbate, M. & Kjelleberg, S. AHL-driven quorum-sensing circuits: their 2414 frequency and function among the Proteobacteria. ISME J. 2, 345 (2008). 2415 13. Pereira, C. S., de Regt, A. K., Brito, P. H., Miller, S. T. & Xavier, K. B. Identification of 2416 functional LsrB-like autoinducer-2 receptors. J. Bacteriol. 191, 6975–6987 (2009). 2417 14. Quan, D. N. & Bentley, W. E. Gene network homology in prokaryotes using a similarity 2418 search approach: Queries of quorum sensing signal transduction. PLoS Comput. Biol. 8, 2419 e1002637 (2012). 2420 15. Bassler, B. L. Small talk: cell-to-cell communication in bacteria. Cell 109, 421–424 (2002). 2421 16. Anetzberger, C. et al. Autoinducers act as biological timers in Vibrio harveyi. PloS One 7, 2422 e48310 (2012). 2423 17. Tu, K. C., Long, T., Svenningsen, S. L., Wingreen, N. S. & Bassler, B. L. Negative 2424 Feedback Loops Involving Small Regulatory RNAs Precisely Control the Vibrio harveyi 2425 Quorum-Sensing Response. Mol. Cell 37, 567–579 (2010). 2426 18. Tu, K. C., Waters, C. M., Svenningsen, S. L. & Bassler, B. L. A small‐RNA‐mediated 2427 negative feedback loop controls quorum‐sensing dynamics in Vibrio harveyi. Mol. 2428 Microbiol. 70, 896–907 (2008). 2429 19. Winson, M. K. et al. Construction and analysis of luxCDABE‐based plasmid sensors for 2430 investigating N‐acyl homoserine lactone‐mediated quorum sensing. FEMS Microbiol. Lett. 2431 163, 185–192 (1998). 2432 20. Swift, S. et al. Quorum sensing in Aeromonas hydrophila and Aeromonas salmonicida: 2433 identification of the LuxRI homologs AhyRI and AsaRI and their cognate N-2434 acylhomoserine lactone signal molecules. J. Bacteriol. 179, 5271–5281 (1997). 2435 134 21. Ng, W.-L. & Bassler, B. L. Bacterial quorum-sensing network architectures. Annu. Rev. 2436 Genet. 43, 197–222 (2009). 2437 22. Waters, C. M. & Bassler, B. L. Quorum sensing: cell-to-cell communication in bacteria. 2438 Annu Rev Cell Dev Biol 21, 319–346 (2005). 2439 23. Cao, J.-G. & Meighen, E. Purification and structural identification of an autoinducer for the 2440 luminescence system of Vibrio harveyi. J. Biol. Chem. 264, 21670–21676 (1989). 2441 24. Milton, D. L. et al. The LuxM Homologue VanM from Vibrio anguillarum directs the 2442 Synthesis of N-(3-Hydroxyhexanoyl) homoserine Lactone and N-Hexanoylhomoserine 2443 Lactone. J. Bacteriol. 183, 3537–3547 (2001). 2444 25. Hanzelka, B. L. et al. Acylhomoserine lactone synthase activity of the Vibrio fischeri AinS 2445 protein. J. Bacteriol. 181, 5766–5770 (1999). 2446 26. Makino, K. et al. Genome sequence of Vibrio parahaemolyticus: a pathogenic mechanism 2447 distinct from that of V cholerae. The Lancet 361, 743–749 (2003). 2448 27. Sun, J., Daniel, R., Wagner-Döbler, I. & Zeng, A.-P. Is autoinducer-2 a universal signal for 2449 interspecies communication: a comparative genomic and phylogenetic analysis of the 2450 synthesis and signal transduction pathways. BMC Evol. Biol. 4, 36 (2004). 2451 28. Schauder, S., Shokat, K., Surette, M. G. & Bassler, B. L. The LuxS family of bacterial 2452 autoinducers: biosynthesis of a novel quorum‐sensing signal molecule. Mol. Microbiol. 41, 2453 463–476 (2001). 2454 29. Chen, X. et al. Structural identification of a bacterial quorum-sensing signal containing 2455 boron. Nature 415, 545–549 (2002). 2456 30. Tait, K., Hutchison, Z., Thompson, F. L. & Munn, C. B. Quorum sensing signal production 2457 and inhibition by coral‐associated Vibrios. Environ. Microbiol. Rep. 2, 145–150 (2010). 2458 31. Ng, W. et al. Signal production and detection specificity in Vibrio CqsA/CqsS quorum‐2459 sensing systems. Mol. Microbiol. 79, 1407–1417 (2011). 2460 32. Gray, K. M. & Garey, J. R. The evolution of bacterial LuxI and LuxR quorum sensing 2461 regulators. Microbiology 147, 2379–2387 (2001). 2462 33. Danino, T., Mondragón-Palomino, O., Tsimring, L. & Hasty, J. A synchronized quorum of 2463 genetic clocks. Nature 463, 326–330 (2010). 2464 34. Bulter, T. et al. Design of artificial cell–cell communication using gene and metabolic 2465 networks. Proc. Natl. Acad. Sci. U. S. A. 101, 2299–2304 (2004). 2466 35. Dai, Y., Toley, B. J., Swofford, C. A. & Forbes, N. S. Construction of an inducible cell‐2467 communication system that amplifies Salmonella gene expression in tumor tissue. 2468 Biotechnol. Bioeng. 110, 1769–1781 (2013). 2469 36. Tamsir, A., Tabor, J. J. & Voigt, C. A. Robust multicellular computing using genetically 2470 encoded NOR gates and chemical/wires/’. Nature 469, 212–215 (2011). 2471 37. Saeidi, N. et al. Engineering microbes to sense and eradicate Pseudomonas aeruginosa, a 2472 human pathogen. Mol. Syst. Biol. 7, (2011). 2473 38. Danino, V. E., Wilkinson, A., Edwards, A. & Downie, J. A. Recipient‐induced transfer of 2474 the symbiotic plasmid pRL1JI in Rhizobium leguminosarum bv. viciae is regulated by a 2475 quorum‐sensing relay. Mol. Microbiol. 50, 511–525 (2003). 2476 39. Ahmer, B. M., van Reeuwijk, J., Timmers, C. D., Valentine, P. J. & Heffron, F. Salmonella 2477 typhimurium encodes an SdiA homolog, a putative quorum sensor of the LuxR family, that 2478 regulates genes on the virulence plasmid. J. Bacteriol. 180, 1185–1193 (1998). 2479 40. Yao, Y. et al. Structure of the Escherichia coli Quorum Sensing Protein SdiA: Activation of 2480 the Folding Switch by Acyl Homoserine Lactones. J. Mol. Biol. 355, 262–273 (2006). 2481 41. Houdt, R., Aertsen, A., Moons, P., Vanoirbeek, K. & Michiels, C. W. N‐acyl‐l‐homoserine 2482 lactone signal interception by Escherichia coli. FEMS Microbiol. Lett. 256, 83–89 (2006). 2483 42. Michael, B., Smith, J. N., Swift, S., Heffron, F. & Ahmer, B. M. SdiA of Salmonella 2484 enterica is a LuxR homolog that detects mixed microbial communities. J. Bacteriol. 183, 2485 5733–5742 (2001). 2486 135 43. Smith, J. N. & Ahmer, B. M. Detection of other microbial species by Salmonella: 2487 expression of the SdiA regulon. J. Bacteriol. 185, 1357–1366 (2003). 2488 44. Sharma, V. K., Bearson, S. M. & Bearson, B. L. Evaluation of the effects of sdiA, a luxR 2489 homologue, on adherence and motility of Escherichia coli O157: H7. Microbiology 156, 2490 1303–1312 (2010). 2491 45. Dyszel, J. L. et al. E. coli K-12 and EHEC genes regulated by SdiA. PLoS One 5, e8946 2492 (2010). 2493 46. Winzer, K. et al. LuxS: its role in central metabolism and the in vitro synthesis of 4-2494 hydroxy-5-methyl-3 (2H)-furanone. Microbiology 148, 909–922 (2002). 2495 47. Bodor, A., Elxnat, B., Thiel, V., Schulz, S. & Wagner-Döbler, I. Potential for luxS related 2496 signalling in marine bacteria and production of autoinducer-2 in the genus Shewanella. 2497 BMC Microbiol. 8, 13 (2008). 2498 48. Tavender, T. J., Halliday, N. M., Hardie, K. R. & Winzer, K. LuxS-independent formation 2499 of AI-2 from ribulose-5-phosphate. BMC Microbiol. 8, 98 (2008). 2500 49. Li, J. et al. A stochastic model of Escherichia coli AI‐2 quorum signal circuit reveals 2501 alternative synthesis pathways. Mol. Syst. Biol. 2, (2006). 2502 50. Meijler, M. M. et al. Synthesis and biological validation of a ubiquitous quorum‐sensing 2503 molecule. Angew. Chem. Int. Ed. 43, 2106–2108 (2004). 2504 51. Vendeville, A., Winzer, K., Heurlier, K., Tang, C. M. & Hardie, K. R. Making’sense’of 2505 metabolism: autoinducer-2, LuxS and pathogenic bacteria. Nat. Rev. Microbiol. 3, 383–396 2506 (2005). 2507 52. Daubin, V., Gouy, M. & Perriere, G. A phylogenomic approach to bacterial phylogeny: 2508 evidence of a core of genes sharing a common history. Genome Res. 12, 1080–1090 (2002). 2509 53. Taga, M. E., Semmelhack, J. L. & Bassler, B. L. The LuxS‐dependent autoinducer AI‐2 2510 controls the expression of an ABC transporter that functions in AI‐2 uptake in Salmonella 2511 typhimurium. Mol. Microbiol. 42, 777–793 (2001). 2512 54. Taga, M. E., Miller, S. T. & Bassler, B. L. Lsr‐mediated transport and processing of AI‐2 in 2513 Salmonella typhimurium. Mol. Microbiol. 50, 1411–1427 (2003). 2514 55. Wang, L., Hashimoto, Y., Tsao, C.-Y., Valdes, J. J. & Bentley, W. E. Cyclic AMP (cAMP) 2515 and cAMP receptor protein influence both synthesis and uptake of extracellular autoinducer 2516 2 in Escherichia coli. J. Bacteriol. 187, 2066–2076 (2005). 2517 56. Wang, L., Li, J., March, J. C., Valdes, J. J. & Bentley, W. E. luxS-dependent gene 2518 regulation in Escherichia coli K-12 revealed by genomic expression profiling. J. Bacteriol. 2519 187, 8350–8360 (2005). 2520 57. Xavier, K. B. & Bassler, B. L. Regulation of uptake and processing of the quorum-sensing 2521 autoinducer AI-2 in Escherichia coli. J. Bacteriol. 187, 238–248 (2005). 2522 58. Herzberg, M., Kaye, I. K., Peti, W. & Wood, T. K. YdgG (TqsA) controls biofilm 2523 formation in Escherichia coli K-12 through autoinducer 2 transport. J. Bacteriol. 188, 587–2524 598 (2006). 2525 59. Rettner, R. E. & Saier Jr, M. H. The autoinducer-2 exporter superfamily. J. Mol. Microbiol. 2526 Biotechnol. 18, 195–205 (2010). 2527 60. Li, J. et al. Quorum sensing in Escherichia coli is signaled by AI-2/LsrR: effects on small 2528 RNA and biofilm architecture. J. Bacteriol. 189, 6011–6020 (2007). 2529 61. Shao, H., James, D., Lamont, R. J. & Demuth, D. R. Differential interaction of 2530 Aggregatibacter (Actinobacillus) actinomycetemcomitans LsrB and RbsB proteins with 2531 autoinducer 2. J. Bacteriol. 189, 5559–5565 (2007). 2532 62. Rader, B. A. et al. Helicobacter pylori perceives the quorum-sensing molecule AI-2 as a 2533 chemorepellent via the chemoreceptor TlpB. Microbiology 157, 2445–2455 (2011). 2534 63. Pereira, C. S. et al. Phosphoenolpyruvate phosphotransferase system regulates detection and 2535 processing of the quorum sensing signal autoinducer‐2. Mol. Microbiol. 84, 93–104 (2012). 2536 136 64. Thompson, J. Galactose transport systems in Streptococcus lactis. J. Bacteriol. 144, 683–2537 691 (1980). 2538 65. Henderson, P. J. Proton-linked sugar transport systems in bacteria. J. Bioenerg. Biomembr. 2539 22, 525–569 (1990). 2540 66. Chaudhuri, B. N., Ko, J., Park, C., Jones, T. A. & Mowbray, S. L. Structure of d-allose 2541 binding protein from Escherichia coli bound to d-allose at 1.8 Å resolution. J. Mol. Biol. 2542 286, 1519–1531 (1999). 2543 67. Shao, H., Lamont, R. J. & Demuth, D. R. Autoinducer 2 is required for biofilm growth of 2544 Aggregatibacter (Actinobacillus) actinomycetemcomitans. Infect. Immun. 75, 4211–4218 2545 (2007). 2546 68. Armbruster, C. E. et al. RbsB (NTHI_0632) mediates quorum signal uptake in nontypeable 2547 Haemophilus influenzae strain 86‐028NP. Mol. Microbiol. 82, 836–850 (2011). 2548 69. Thijs, I. M. et al. The AI-2-dependent regulator LsrR has a limited regulon in Salmonella 2549 Typhimurium. Cell Res. 20, 966–969 (2010). 2550 70. Byrd, C. M. Local and global gene regulation analysis of the autoinducer-2 mediated 2551 quorum sensing mechanism in Escherichia coli. (2011). 2552 71. Pesavento, C. et al. Inverse regulatory coordination of motility and curli-mediated adhesion 2553 in Escherichia coli. Genes Dev. 22, 2434–2446 (2008). 2554 72. Park, J. T., Raychaudhuri, D., Li, H., Normark, S. & Mengin-Lecreulx, D. MppA, a 2555 Periplasmic Binding Protein Essential for Import of the Bacterial Cell Wall Peptidel-2556 Alanyl-γ-d-Glutamyl-meso-Diaminopimelate. J. Bacteriol. 180, 1215–1223 (1998). 2557 73. Diaz, Z., Xavier, K. B. & Miller, S. T. The crystal structure of the Escherichia coli 2558 autoinducer-2 processing protein LsrF. PloS One 4, e6820 (2009). 2559 74. Marques, J. C. et al. Processing the Interspecies Quorum-sensing Signal Autoinducer-2 (AI-2560 2) Characterization phospho-(S)-4, 5-dihydroxy-2, 3-pentanedione isomerization by LsrG 2561 protein. J. Biol. Chem. 286, 18331–18343 (2011). 2562 75. Xavier, K. B. et al. Phosphorylation and processing of the quorum-sensing molecule 2563 autoinducer-2 in enteric bacteria. ACS Chem. Biol. 2, 128–136 (2007). 2564 76. Pellicer, M. T., Nuñez, M. F., Aguilar, J., Badia, J. & Baldoma, L. Role of 2-2565 phosphoglycolate phosphatase of Escherichia coli in metabolism of the 2-phosphoglycolate 2566 formed in DNA repair. J. Bacteriol. 185, 5815–5821 (2003). 2567 77. Choi, J. et al. LsrR-mediated quorum sensing controls invasiveness of Salmonella 2568 typhimurium by regulating SPI-1 and flagella genes. PloS One 7, e37059 (2012). 2569 78. Niu, C. et al. LuxS influences Escherichia coli biofilm formation through autoinducer‐2‐2570 dependent and autoinducer‐2‐independent modalities. FEMS Microbiol. Ecol. 83, 778–791 2571 (2013). 2572 79. Torres-Escobar, A., Juárez-Rodríguez, M. D., Lamont, R. J. & Demuth, D. R. 2573 Transcriptional regulation of Aggregatibacter actinomycetemcomitans lsrACDBFG and 2574 lsrRK operons and their role in biofilm formation. J. Bacteriol. 195, 56–65 (2013). 2575 80. Darch, S. E., West, S. A., Winzer, K. & Diggle, S. P. Density-dependent fitness benefits in 2576 quorum-sensing bacterial populations. Proc. Natl. Acad. Sci. 109, 8259–8263 (2012). 2577 81. Krin, E. et al. Pleiotropic role of quorum-sensing autoinducer 2 in Photorhabdus 2578 luminescens. Appl. Environ. Microbiol. 72, 6439–6451 (2006). 2579 82. Auger, S., Krin, E., Aymerich, S. & Gohar, M. Autoinducer 2 affects biofilm formation by 2580 Bacillus cereus. Appl. Environ. Microbiol. 72, 937–941 (2006). 2581 83. Li, L. et al. Analysis on Actinobacillus pleuropneumoniae LuxS regulated genes reveals 2582 pleiotropic roles of LuxS/AI-2 on biofilm formation, adhesion ability and iron metabolism. 2583 Microb. Pathog. 50, 293–302 (2011). 2584 84. Vidal, J. E., Ludewick, H. P., Kunkel, R. M., Zähner, D. & Klugman, K. P. The LuxS-2585 dependent quorum-sensing system regulates early biofilm formation by Streptococcus 2586 pneumoniae strain D39. Infect. Immun. 79, 4050–4060 (2011). 2587 137 85. Li, M., Villaruz, A. E., Vadyvaloo, V., Sturdevant, D. E. & Otto, M. AI-2-dependent gene 2588 regulation in Staphylococcus epidermidis. BMC Microbiol. 8, 4 (2008). 2589 86. Ahmed, N. A., Petersen, F. C. & Scheie, A. A. AI-2/LuxS is involved in increased biofilm 2590 formation by Streptococcus intermedius in the presence of antibiotics. Antimicrob. Agents 2591 Chemother. 53, 4258–4263 (2009). 2592 87. Von Lackum, K. et al. Functionality of Borrelia burgdorferi LuxS: The Lyme disease 2593 spirochete produces and responds to the pheromone autoinducer-2 and lacks a complete 2594 activated-methyl cycle. Int. J. Med. Microbiol. 296, 92–102 (2006). 2595 88. Rader, B. A., Campagna, S. R., Semmelhack, M. F., Bassler, B. L. & Guillemin, K. The 2596 quorum-sensing molecule autoinducer 2 regulates motility and flagellar morphogenesis in 2597 Helicobacter pylori. J. Bacteriol. 189, 6109–6117 (2007). 2598 89. Pereira, C. S., McAuley, J. R., Taga, M. E., Xavier, K. B. & Miller, S. T. Sinorhizobium 2599 meliloti, a bacterium lacking the autoinducer‐2 (AI‐2) synthase, responds to AI‐2 supplied 2600 by other bacteria. Mol. Microbiol. 70, 1223–1235 (2008). 2601 90. Duan, K., Dammel, C., Stein, J., Rabin, H. & Surette, M. G. Modulation of Pseudomonas 2602 aeruginosa gene expression by host microflora through interspecies communication. Mol. 2603 Microbiol. 50, 1477–1491 (2003). 2604 91. Geier, H., Mostowy, S., Cangelosi, G. A., Behr, M. A. & Ford, T. E. Autoinducer-2 triggers 2605 the oxidative stress response in Mycobacterium avium, leading to biofilm formation. Appl. 2606 Environ. Microbiol. 74, 1798–1804 (2008). 2607 92. Strassmann, J. E., Gilbert, O. M. & Queller, D. C. Kin discrimination and cooperation in 2608 microbes. Annu. Rev. Microbiol. 65, 349–367 (2011). 2609 93. Wu, K. & Rao, C. V. The role of configuration and coupling in autoregulatory gene circuits. 2610 Mol. Microbiol. 75, 513–527 (2010). 2611 94. Tsao, C.-Y., Hooshangi, S., Wu, H.-C., Valdes, J. J. & Bentley, W. E. Autonomous 2612 induction of recombinant proteins by minimally rewiring native quorum sensing regulon of 2613 E. coli. Metab. Eng. 12, 291–297 (2010). 2614 95. Gonzalez Barrios, A. F. & Achenie, L. E. Escherichia coli autoinducer-2 uptake network 2615 does not display hysteretic behavior but AI-2 synthesis rate controls transient bifurcation. 2616 Biosystems 99, 17–26 (2010). 2617 96. Englert, D. L., Manson, M. D. & Jayaraman, A. Flow-based microfluidic device for 2618 quantifying bacterial chemotaxis in stable, competing gradients. Appl. Environ. Microbiol. 2619 75, 4557–4564 (2009). 2620 97. Hegde, M. et al. Chemotaxis to the quorum-sensing signal AI-2 requires the Tsr 2621 chemoreceptor and the periplasmic LsrB AI-2-binding protein. J. Bacteriol. 193, 768–773 2622 (2011). 2623 98. Zahedmanesh, H. & Lally, C. A multiscale mechanobiological modelling framework using 2624 agent-based models and finite element analysis: application to vascular tissue engineering. 2625 Biomech. Model. Mechanobiol. 11, 363–377 (2012). 2626 99. Suyama, M. & Bork, P. Evolution of prokaryotic gene order: genome rearrangements in 2627 closely related species. Trends Genet. 17, 10–13 (2001). 2628 100. Moreno-Hagelsieb, G. & Collado-Vides, J. A powerful non-homology method for the 2629 prediction of operons in prokaryotes. Bioinformatics 18, S329–S336 (2002). 2630 101. Overbeek, R., Fonstein, M., D’souza, M., Pusch, G. D. & Maltsev, N. The use of gene 2631 clusters to infer functional coupling. Proc. Natl. Acad. Sci. 96, 2896–2901 (1999). 2632 102. Touchon, M. et al. Organised genome dynamics in the Escherichia coli species results in 2633 highly diverse adaptive paths. PLoS Genet. 5, e1000344 (2009). 2634 103. Fang, G., Rocha, E. P. & Danchin, A. Persistence drives gene clustering in bacterial 2635 genomes. BMC Genomics 9, 4 (2008). 2636 104. Harrington, E. D., Jensen, L. J. & Bork, P. Predicting biological networks from genomic 2637 data. FEBS Lett. 582, 1251–1258 (2008). 2638 138 105. Rogozin, I. B., Makarova, K. S., Wolf, Y. I. & Koonin, E. V. Computational approaches for 2639 the analysis of gene neighbourhoods in prokaryotic genomes. Brief. Bioinform. 5, 131–149 2640 (2004). 2641 106. Altschul, S. F., Gish, W., Miller, W., Myers, E. W. & Lipman, D. J. Basic local alignment 2642 search tool. J. Mol. Biol. 215, 403–410 (1990). 2643 107. Stajich, J. E. et al. The Bioperl toolkit: Perl modules for the life sciences. Genome Res. 12, 2644 1611–1618 (2002). 2645 108. Williams, K. P. et al. Phylogeny of gammaproteobacteria. J. Bacteriol. 192, 2305–2314 2646 (2010). 2647 109. Roderick, S. L. The lac operon galactoside acetyltransferase. C. R. Biol. 328, 568–575 2648 (2005). 2649 110. Arraj, J. A. & Campbell, J. H. Isolation and characterization of the newly evolved ebg beta-2650 galactosidase of Escherichia coli K-12. J. Bacteriol. 124, 849–856 (1975). 2651 111. Stoebel, D. M. Lack of evidence for horizontal transfer of the lac operon into Escherichia 2652 coli. Mol. Biol. Evol. 22, 683–690 (2005). 2653 112. Hooshangi, S. & Bentley, W. E. LsrR quorum sensing ‘switch’ is revealed by a bottom-up 2654 approach. PLoS Comput. Biol. 7, e1002172 (2011). 2655 113. Studier, F. W., Daegelen, P., Lenski, R. E., Maslov, S. & Kim, J. F. Understanding the 2656 Differences between Genome Sequences of Escherichia coli B Strains REL606 and BL21 2657 (DE3) and Comparison of the E. coli B and K-12 Genomes. J. Mol. Biol. 394, 653–680 2658 (2009). 2659 114. Xue, T., Zhao, L., Sun, H., Zhou, X. & Sun, B. LsrR-binding site recognition and regulatory 2660 characteristics in Escherichia coli AI-2 quorum sensing. Cell Res. 19, 1258–1268 (2009). 2661 115. Price, M. N., Dehal, P. S. & Arkin, A. P. Horizontal gene transfer and the evolution of 2662 transcriptional regulation in Escherichia coli. Genome Biol 9, R4 (2008). 2663 116. Mavromatis, K. et al. Complete genome sequence of Spirochaeta smaragdinae type strain 2664 (SEBR 4228T). Stand. Genomic Sci. 3, 136 (2010). 2665 117. Tanouchi, Y., Tu, D., Kim, J. & You, L. Noise reduction by diffusional dissipation in a 2666 minimal quorum sensing motif. PLoS Comput. Biol. 4, e1000167 (2008). 2667 118. Tabareau, N., Slotine, J.-J. & Pham, Q.-C. How synchronization protects from noise. PLoS 2668 Comput. Biol. 6, e1000637 (2010). 2669 119. Anetzberger, C., Pirch, T. & Jung, K. Heterogeneity in quorum sensing‐regulated 2670 bioluminescence of Vibrio harveyi. Mol. Microbiol. 73, 267–277 (2009). 2671 120. Sayut, D. J., Kambam, P. K. R. & Sun, L. Noise and kinetics of LuxR positive feedback 2672 loops. Biochem. Biophys. Res. Commun. 363, 667–673 (2007). 2673 121. Kittisopikul, M. & Süel, G. M. Biological role of noise encoded in a genetic network motif. 2674 Proc. Natl. Acad. Sci. 107, 13300–13305 (2010). 2675 122. Fenley, A. T., Banik, S. K. & Kulkarni, R. V. Computational modeling of differences in the 2676 quorum sensing induced luminescence phenotypes of Vibrio harveyi and Vibrio cholerae. J. 2677 Theor. Biol. 274, 145–153 (2011). 2678 123. Marques, J. C. et al. LsrF, a coenzyme A-dependent thiolase, catalyzes the terminal step in 2679 processing the quorum sensing signal autoinducer-2. Proc. Natl. Acad. Sci. 111, 14235–2680 14240 (2014). 2681 124. Parsek, M. R. & Greenberg, E. Sociomicrobiology: the connections between quorum 2682 sensing and biofilms. Trends Microbiol. 13, 27–33 (2005). 2683 125. Defoirdt, T. Can bacteria actively search to join groups? ISME J.-Int. Soc. Microb. Ecol. 5, 2684 569 (2011). 2685 126. Sourjik, V. & Armitage, J. P. Spatial organization in bacterial chemotaxis. EMBO J. 29, 2686 2724–2733 (2010). 2687 127. Dockery, J. D. & Keener, J. P. A mathematical model for quorum sensing in Pseudomonas 2688 aeruginosa. Bull. Math. Biol. 63, 95–116 (2001). 2689 139 128. Bentley, W. E., Mirjalili, N., Andersen, D. C., Davis, R. H. & Kompala, D. S. Plasmid‐2690 encoded protein: the principal factor in the ‘metabolic burden’ associated with recombinant 2691 bacteria. Biotechnol. Bioeng. 35, 668–681 (1990). 2692 129. Melke, P., Sahlin, P., Levchenko, A. & Jönsson, H. A cell-based model for quorum sensing 2693 in heterogeneous bacterial colonies. PLoS Comput. Biol. 6, e1000819 (2010). 2694 130. Mittal, N., Budrene, E. O., Brenner, M. P. & van Oudenaarden, A. Motility of Escherichia 2695 coli cells in clusters formed by chemotactic aggregation. Proc. Natl. Acad. Sci. 100, 13259–2696 13263 (2003). 2697 131. Wu, H. et al. Autonomous bacterial localization and gene expression based on nearby cell 2698 receptor density. Mol. Syst. Biol. 9, (2013). 2699 132. Albert, R. & Othmer, H. G. The topology of the regulatory interactions predicts the 2700 expression pattern of the segment polarity genes in Drosophila melanogaster. J. Theor. 2701 Biol. 223, 1–18 (2003). 2702 133. Lawrence, P. A., Casal, J. & Struhl, G. hedgehog and engrailed: pattern formation and 2703 polarity in the Drosophila abdomen. Development 126, 2431–2439 (1999). 2704 134. Birtwistle, M. R. et al. Emergence of bimodal cell population responses from the interplay 2705 between analog single-cell signaling and protein expression noise. BMC Syst. Biol. 6, 109 2706 (2012). 2707 135. Shalek, A. K. et al. Single-cell transcriptomics reveals bimodality in expression and 2708 splicing in immune cells. Nature (2013). 2709 136. Veening, J.-W., Smits, W. K. & Kuipers, O. P. Bistability, epigenetics, and bet-hedging in 2710 bacteria. Annu Rev Microbiol 62, 193–210 (2008). 2711 137. De Beer, D. & Stoodley, P. Relation between the structure of an aerobic biofilm and 2712 transport phenomena. Water Sci. Technol. 32, 11–18 (1995). 2713 138. Coulthurst, S. J., Kurz, C. L. & Salmond, G. P. luxS mutants of Serratia defective in 2714 autoinducer-2-dependent ‘quorum sensing’show strain-dependent impacts on virulence and 2715 production of carbapenem and prodigiosin. Microbiology 150, 1901–1910 (2004). 2716 139. Høyland-Kroghsbo, N. M., Mærkedahl, R. B. & Svenningsen, S. L. A quorum-sensing-2717 induced bacteriophage defense mechanism. MBio 4, e00362–12 (2013). 2718 140. Sun, S., Kjelleberg, S. & McDougald, D. Relative contributions of Vibrio polysaccharide 2719 and quorum sensing to the resistance of Vibrio cholerae to predation by heterotrophic 2720 protists. PloS One 8, e56338 (2013). 2721 141. Diggle, S. P., Griffin, A. S., Campbell, G. S. & West, S. A. Cooperation and conflict in 2722 quorum-sensing bacterial populations. Nature 450, 411–414 (2007). 2723 142. Dandekar, A. A., Chugani, S. & Greenberg, E. P. Bacterial quorum sensing and metabolic 2724 incentives to cooperate. Science 338, 264–266 (2012). 2725 143. Byrd, C. M. & Bentley, W. E. Quieting cross talk - the quorum sensing regulator LsrR as a 2726 possible target for fighting bacterial infections. Cell Res 19, 1229–1230 2727 144. Luo, X. et al. Biofabrication of stratified biofilm mimics for observation and control of 2728 bacterial signaling. Biomaterials 33, 5136–5143 (2012). 2729 145. Goryachev, A. B. Understanding Bacterial Cell− Cell Communication with Computational 2730 Modeling. Chem Rev 111, 238–250 (2011). 2731 146. Wilke, C. & Chang, P. Correlation of diffusion coefficients in dilute solutions. AIChE J. 1, 2732 264–270 (1955). 2733 147. Terrell, J. L. et al. Integrated biofabrication for electro‐addressed in‐film bioprocessing. 2734 Biotechnol. J. 7, 428–439 (2012). 2735 148. Becskei, A., Séraphin, B. & Serrano, L. Positive feedback in eukaryotic gene networks: cell 2736 differentiation by graded to binary response conversion. EMBO J. 20, 2528–2535 (2001). 2737 149. Ozbudak, E. M., Thattai, M., Lim, H. N., Shraiman, B. I. & Van Oudenaarden, A. 2738 Multistability in the lactose utilization network of Escherichia coli. Nature 427, 737–740 2739 (2004). 2740 140 150. Zhu, J. & Pei, D. A LuxP-based fluorescent sensor for bacterial autoinducer II. ACS Chem. 2741 Biol. 3, 110–119 (2008). 2742 151. Wu, M., Tao, Y., Liu, X. & Zang, J. Structural basis for phosphorylated autoinducer-2 2743 modulation of the oligomerization state of the global transcription regulator LsrR from 2744 Escherichia coli. J. Biol. Chem. 288, 15878–15887 (2013). 2745 152. Ganguly, A., Rajdev, P., Williams, S. M. & Chatterji, D. Nonspecific Interaction between 2746 DNA and Protein allows for Cooperativity: A Case Study with Mycobacterium DNA 2747 Binding Protein. J. Phys. Chem. B 116, 621–632 (2011). 2748 153. Von Dassow, G., Meir, E., Munro, E. M. & Odell, G. M. The segment polarity network is a 2749 robust developmental module. Nature 406, 188–192 (2000). 2750 154. Matsumoto, M. & Nishimura, T. Mersenne twister: a 623-dimensionally equidistributed 2751 uniform pseudo-random number generator. ACM Trans. Model. Comput. Simul. TOMACS 2752 8, 3–30 (1998). 2753 2754