Possible Drivers in Endophyte Diversity and Transmission in the Plant Bacterial

Microbiome

Thesis

Presented in Partial Fulfillment of the Requirements for the Degree Master of Science in

the Graduate School of The Ohio State University

By

Ana María Vázquez, B.S.

Graduate Program in Plant Pathology

The Ohio State University

2020

Thesis Committee

Dr. María Soledad Benítez-Ponce, Advisor

Dr. Christine Sprunger

Dr. Jonathan M. Jacobs

1

Copyrighted by

Ana María Vázquez

2020

2

Abstract

It has been documented that beneficial plant-associated have contributed to disease suppression, growth promotion, and tolerance to abiotic stresses. Advances in high-throughput sequencing have allowed an increase in research regarding bacterial endophytes, which are microbes that colonize the interior of plants without causing disease. Practices associated with minimizing the use of off-farm resources, such as reduced tillage regimes and crop rotations, can cause shifts in plant-associated bacteria and its surrounding agroecosystem. Integrated crop–livestock systems are an option that can provide environmental benefits by implementing diverse cropping systems, incorporating perennial and legume forages and adding animal manure through grazing livestock. It has been found that crop-livestock systems can increase soil quality and fertility, reduce cost of herbicide use and improve sustainability, especially for farmers in poorer areas of the world. This work explores how crop-livestock systems that integrate chicken rotations can impact tomato plant growth, as well as soil and endophytic bacterial communities. Tomato plants were subjected to greenhouse and field studies where biomass was assessed, and bacterial communities were characterized through culture- dependent and -independent approaches.

ii

In greenhouse experiments, the greater percent of chicken grazed soil incorporated in the planting substrate, the greater the stunting of tomato seedlings. In the field study, bacterial communities differed significantly by sample origin and plant development stage, regardless of chicken grazing history. Our findings suggest stronger contribution of agricultural management practices during early plant stages on endophytic microbiome, as opposed to later on in the host lifecycle. Taxonomic composition of dominant groups of recovered endophytic bacterial isolates were consistent with those found by amplicon sequencing. Plots with history of chicken grazing had a significant increase of soil fertility and differentially abundant ASVs. Dominant phyla

(, , and ) of bacterial taxa identified were uniform between culture-dependent and -independent approaches from both greenhouse and field experiments. Most of the recovered isolates were found to be phylogenetically similar to formerly cultured plant endophytes and bacteria found in soil.

However, some isolates were similar to bacteria recovered in human clinical settings.

Further directions should focus on identification and elucidation of transmission of possible plant growth promoting bacteria, as well as human pathogens, found in crops grown in this type of agricultural management strategy. Moreover, the influence of stage in plant lifecycle on bacterial community composition under specific management strategies remains to be explored. This work provides insight into the influence of crop- livestock rotations on soil- and tomato plant-associated bacterial communities.

iii

Dedication

This thesis is dedicated to the memory of my biggest fan, my grandmother

Carmen Leticia Cabrera Amador. Regardless of adversities, Abuela lived by seeing beauty in science, art and human connections. I will always honor her by emulating her tenacity and persistence.

iv

Acknowledgments

First, I would like to acknowledge my advisor, María Soledad Benítez-Ponce for her support and commitment in mentoring me during these past two years. Thank you for always stimulating an intellectually challenging and collaborative environment where I could thrive as a scientist. In addition, thank you for encouraging me to integrate other aspects to my professional skillset.

Moreover, I would like to thank the members of my student advisory committee,

Christine Sprunger and Jonathan Jacobs, for their valuable and constructive suggestions. I would also like to thank Suranga Basnagala for his advice in setting up my first field experiment at Mellinger Research Farm, as well as his assistance in looking out for my tomato plants and answering all of my questions.

I would like to offer my special thanks to Leslie Taylor for all her help with my experiments and much needed cheer during this journey. Through countless emails exchanged, assistance with computational analysis provided by former lab member

Matthew Willman was greatly appreciated. In addition, I would like to thank lab members and fellow graduate students Daowen Huo and Ranjana Rawal for their inspiring motivation and persistence. A special thanks to the whole Benitez lab, current v and former members, for making my time in Wooster memorable and cherished. Thank you to undergraduates Nedas Matulionis, Emmily Moses and Yesenia Velez for helping in different aspects of my research project, in and outside the lab.

My grateful thanks are also extended to three very special people that have helped me make Wooster a home away from home, Marlia Bosques, Madeline Horvat and

Edwin Navarro. In addition, a big thanks to all my friends, back home and in different places, that have encouraged me from afar.

Finally, the biggest thanks are for the first supporters and mentors I ever had, my family. Mamá, Michelle, Berto, Carlos, José, Tommy and Tito. Thanks to my abuela for being my biggest and, arguably, my most outspoken fan.

vi

Vita

1995………………………………………… Born, Arecibo, Puerto Rico

2013-2018……………………………………B.S. Biology and B.S. Industrial

Microbiology, University of Puerto Rico at Mayagüez

2018 to present………………………………Graduate Research Associate, Department of Plant Pathology, The Ohio State University

Fields of Study

Major Field: Plant Pathology

vii

Table of Contents

Abstract ...... ii

Dedication ...... iv

Acknowledgments ...... v

Vita ...... vii

List of Tables ...... xii

List of Figures ...... xvi

Chapter 1. Literature Review ...... 1

Introduction ...... 1

The Plant Microbiome ...... 3

Bacterial Endophytes in Aboveground Plant Tissues ...... 5

Bacteria in Soil and Rhizosphere of Plants ...... 6

Transmission and Colonization of Plant-Associated Bacterial Communities ...... 8

Role of Bacterial Endophytes ...... 12

Summary and Conclusions ...... 15

viii

References ...... 18

Chapter 2. Characterization of the Bacterial Endophytic Microbiome of Tomato

Seedlings Grown in a Soil Diversity Gradient Using Chicken-Grazed Soils ...... 35

Introduction ...... 35

Materials and Methods ...... 40

Field Site Description ...... 40

Soil Sampling ...... 41

Planting Substrate, Plant Materials and Growth Conditions ...... 42

Seedling Sampling and Evaluation ...... 44

Plant Tissue Sampling and Isolation of Endophytic Bacteria from Tomato Seedling

Stem ...... 45

DNA Extraction, PCR Amplification and Sequencing of the 16S rRNA and rpoB

Gene Regions from Individual Bacterial Isolates ...... 45

Isolate Sequence Data Analysis ...... 48

Plant Growth Data Analysis and Statistics ...... 49

Results ...... 50

Effect of Soil Dilution Treatments on Aboveground Biomass and Height of Tomato

Seedlings ...... 50

Effect of Soil Slurry Dilution Treatments on Aboveground Biomass, Root Biomass,

Height and Number of Compound Leaves of Tomato Seedlings ...... 50 ix

Impact of Dilution Treatments on Cultured Endophytic Bacteria Composition ...... 51

Discussion ...... 53

Tables ...... 82

Figures...... 93

Chapter 3. Characterization of the Endophytic Bacterial Microbiome of Tomato Plants

Grown in a Chicken-Grazed Plots ...... 100

Introduction ...... 100

Materials and Methods ...... 104

Field Site Description ...... 104

Preparation of Tomato Seedlings for Transplant ...... 105

Soil Health Measurements ...... 106

Tomato Plant Sampling and Evaluation ...... 107

Plant Tissue Sampling, Surface Sterilization and Isolation of Endophytic Bacteria

from Tomato Plant Stem ...... 108

Cultured Endophyte DNA Extraction, PCR Amplification and Sequencing of the

16S rRNA Gene Region ...... 109

Isolate Sequence Data Analysis ...... 111

Culture-Independent Characterization of Bacterial Endophytes ...... 111

Amplicon Sequence Data Processing and Analysis ...... 113

x

Amplicon Data Analysis and Statistics ...... 114

Plant Growth and Soil Health Data Analysis and Statistics ...... 116

Results ...... 117

Impact of Chicken Grazing History on Soil Health ...... 117

Effect of Chicken Grazing History on Plant Tissue Measurements of Tomato Plants

Sampled in July and September ...... 117

Impact of Chicken Grazing History on Cultured Endophytic Bacteria Composition

...... 118

Tomato-Associated Bacterial Communities and Diversity Analyses ...... 119

Discussion ...... 123

References ...... 136

Tables ...... 154

Figures...... 174

Bibliography ...... 192

Appendix A. Supplementary Table ...... 225

xi

List of Tables

Table 2.1. History of management strategy rotations implemented in plots selected for soil sampling of 2018 and 2020 greenhouse experiments. For 2018 and 2019 plots sampled are indicated in bold. P/C: Pasture/chickens. P/V: Pasture/vegetables...... 82

Table 2.2. Impact of dilution of chicken grazed soils on aboveground biomass and height, as estimated by analysis of variance, of Celebrity and Moneymaker tomato seedlings.

Seedlings were grown under greenhouse conditions in chicken-grazed soils diluted with a sterile soil substrate in 2018...... 83

Table 2.3. Number of CFU/0.1mL for each chicken-grazed soil slurry dilution. The first inoculation of soil slurry was done prior to tomato seed germination. The second inoculation was added to germinated seedlings...... 84

Table 2.4. Impact of dilution of chicken grazed soils on aboveground biomass, root biomass, height, number of compound leaves and root to shoot ratio of tomato seedlings grown under greenhouse conditions in sterile sand-turface and composted soil substrate with an addition of diluted chicken-grazed soil slurry, for 2020 experiment...... 85

Table 2.5. Soil dilution treatment and block effect on aboveground biomass, root biomass, height, number of compound leaves and root to shoot ratio of tomato seedlings grown under greenhouse conditions in sterile sand-turface and composted soil substrate

xii with an addition of diluted chicken-grazed soil slurry. Means in a column without a common superscript letter differ (p < 0.05), as analyzed by Fisher’s LSD test...... 86

Table 2.6. Bacterial endophytes recovered from the stem of tomato seedlings grown in chicken-grazed soil dilutions. Identification is based on partial 16S sequence. The number of isolates recovered per each sequence type is shown for each treatment and tomato genotype. Abbreviations: CGS: Chicken-grazed soil, SSS: Sterile soil substrate,

C: Celebrity, M: Moneymaker...... 87

Table 2.7. Subset of bacterial endophytes recovered from the stem of tomato seedlings grown in chicken-grazed soil dilutions. Identification is based in rpoB gene region...... 90

Table 2.8. Number of isolates recovered from surface sterilized stem of tomato seedlings grown across all chicken-grazed soil slurry dilutions...... 92

Table 3.1. Effect of chicken grazing history on active pools of organic matter in soil collected from plots prior to planting of the tomato...... 154

Table 3.2. Impact of block and chicken-grazing history on aboveground biomass and number of compound leaves of tomato plants sampled in July...... 155

Table 3.3. Impact of block and chicken-grazing history on aboveground biomass and number of compound leaves of tomato plants sampled in July. Means in a column without a common superscript letter differ (p < 0.05), as analyzed by Fisher’s LSD test.

...... 156

Table 3.4. Impact of block and chicken-grazing history on above-ground biomass, number of compound leaves, total biomass of tomatoes harvested, number of

xiii symptomatic tomatoes harvested, number of total tomatoes harvested and root biomass of tomato plants sampled in September...... 157

Table 3.5. Impact of block and chicken-grazing history on aboveground biomass, number of compound leaves, total biomass of tomatoes harvested, number of symptomatic tomatoes, number of total tomatoes harvested and root biomass of tomato plants sampled in September. Means in a row without a common superscript letter differ (p < 0.1), as analyzed by Fisher’s LSD test...... 159

Table 3.6. Identification and distribution of bacterial endophytes recovered from the stem and tomato flesh of tomato plants grown in plots with and without history of chicken grazing. Abbreviations: J: July, SP: September, S: Stem, TF: Tomato flesh, CG: Chicken grazed, PO: Pasture only...... 160

Table 3.7. Recovery of amplicon sequencing reads from different tomato plant tissues, rhizosphere soil and baseline samples (bulk soil and chicken manure). Using Cutadapt software, each primer and adapter was removed...... 163

Table 3.8. Recovery of amplicon sequencing reads from different tomato plant tissues, rhizosphere soil and baseline samples (bulk soil and chicken manure). Using DADA2, reads were processed and filtered, per sample sequenced...... 164

Table 3.9. Recovery of amplicon sequencing reads from different tomato plant tissues, rhizosphere soil and baseline samples (bulk soil and chicken manure). Using phyloseq, reads belonging to unwanted taxa were removed...... 165

Table 3.10. Increase in bacterial community richness, evenness and diversity in samples obtained from chicken-grazed pasture plots. Alpha diversity metrics (Chao1

xiv richness, Shannon’s diversity index and Faith’s Phylogenetic diversity) of bacterial communities from bulk soil and chicken manure collected in May and September, respectively...... 166

Table 3.11. Variations in bacterial community richness, evenness and diversity in samples obtained from chicken-grazed pasture plots and pasture only plots. Alpha diversity metrics (Chao1 species richness, Shannon’s diversity index and Faith’s

Phylogenetic diversity) of bacterial communities from rhizosphere soil, root, stem and leaf of tomato plants sampled in June and Septem ...... 167

Table 3.12. Variations in bacterial community richness, evenness and diversity in tomato fruit and seed samples obtained from chicken-grazed pasture plots and pasture only plots.

Alpha diversity metrics (Chao1 species richness, Shannon’s diversity index and Faith’s

Phylogenetic diversity) of bacterial communities from tomato flesh and seed sampled in

September...... 168

Table 3.13. Alpha diversity metrics (Chao1 species richness, Shannon’s diversity index and Faith’s Phylogenetic diversity) of bacterial communities from rhizosphere soil and leaf from both sampling months and treatments...... 169

Table 3.14. Differentially abundant taxa (p<0.1) between chicken grazed pasture plots and pasture only plots within sampling months. All taxa shown, except (indicate which) were more abundant in chicken grazed plots...... 171

Table A.1. Identification of differentially abundant taxa (p<0.1) by comparing chicken grazed pasture plots with pasture only plots...... 226

xv

List of Figures

Figure 2.1. Field layout of the SMSFG plots at Mellinger Research Farm at OARDC,

Wooster, Ohio. The 0.6 hectare field is divided in four replicates of five plot treatments installed west of the farmstead...... 93

Figure 2.2. Gradient of diluted soil from chicken-grazed pasture plots. Five treatments were generated by diluting soil from chicken-grazed pasture plots with autoclaved soil substrate: 100% chicken-grazed soil, 75%, 50% and 25% dilutions and 100% autoclaved soil substrate...... 94

Figure 2.3. Gradient of filtered chicken-grazed soil slurry dilution. Filtered soil slurry was diluted to 10-1, 10-2, 10-3 and 10-4. Sterile water was used as a control...... 95

Figure 2.4. Differences in aboveground biomass and height of tomato seedlings grown in chicken-grazed soil dilutions after six weeks (n=5, except for Celebrity 75% chicken- grazed soil n=4). A. Aboveground biomass. B. Height. Celebrity and Moneymaker seeds were sowed in a dilution series of chicken-grazed soils using sterile soil substrate.

Replicates of each treatment were placed in the greenhouse for six weeks and these were measured before being destructively sampled. ANOVA and Fisher’s LSD tests were calculated, and replicate measurements were graphed as boxplots. Black horizontal lines indicate the median, the lower portion of the box is the first quartile, the upper portion of the box is the third quartile, the whiskers are maximum and minimum, and the outliers xvi are black dots. Uppercase letters indicate significant difference across soil dilutions according to Fisher’s LSD test (p < 0.05) for the Celebrity seedlings and the lowercase letters for Moneymaker seedlings. 2018 experiment...... 96

Figure 2.5. Total number of bacterial endophytes recovered by genera. Pie chart indicating the percentage recovery of different bacterial genera from tomato seedling stems grown in different soil dilutions (total = 87 isolates)...... 97

Figure 2.6. Impact of dilution of chicken grazed soils on taxa of bacterial endophytes recovered from two tomato varieties, grown in different soil dilution treatments. The bar plot shows the number of bacterial isolates recovered from tomato seedling stems (total =

87 isolates). Different colors indicate the taxa to which each isolate belongs...... 98

Figure 2.7. Phylogenetic reconstruction of unique endophytic bacterial taxa recovered from the stem of tomato seedlings based on 16S rRNA gene sequences. Taxa labeled as

AVC represent unique taxa recovered in this study and are color coded according to their treatment origin. Reference sequences obtained from the NCBI Database are labeled according to the sample origin. The tree was created using the Maximum Likelihood method based on the Tamura-Nei model and generated in MEGA. The bootstrap consensus tree inferred is from 1000 replicates and branches corresponding to partitions reproduced in less than 60% bootstrap replicates are collapsed. Ferroplasma acidiphilum and Methanobacter cuticularis were used as outgroups...... 99

Figure 3.1. Overview of the field trial conducted in the Summer of 2019 at Mellinger

Research Farm. Plots grazed by chickens in August 2017 and 2018 were used as

xvii vegetable plots in 2019. Tomato plants were grown in soils with differing chicken grazing history...... 174

Figure 3.2. Overview of the three sampling months, involving the bulk soil, tomato plant and chicken manure sampling. At each sampling time, different samples were collected and analyzed for bacterial diversity, plant vigor and soil characteristics...... 175

Figure 3.3. Soil sampling and homogenization for subsequent use in amplicon sequencing, as planting substrate and soil health tests. Soil obtained from chicken-grazed pasture plots and pasture only plots at Mellinger Research Farm. Soil was bagged in separate plastic bags per split plot. In the greenhouse, soil was poured into plastic containers, separated by treatment and plot, and mixed-together and homogenized by hand...... 176

Figure 3.4. Three-week old tomato seedling transplant. On June 14th, 10 tomato seedlings were selected per treatment and plot to be transplanted to the corresponding split plot were soil was originally collected from...... 177

Figure 3.5. Amplicon sequence data processing and analysis pipeline adapted from Huo,

2020...... 178

Figure 3.6. Endophytic bacterial isolates recovered from stem tissue of tomato plants grown in plots with or without history of chicken grazing. Different colors indicate the taxa to which each isolate belongs...... 179

Figure 3.7. Endophytic bacterial isolates recovered from tomato flesh of tomatoes harvested from plants grown in plots with or without history of chicken grazing. Different colors indicate the taxa to which each isolate belongs...... 180

xviii

Figure 3.8. Sequencing depth for each sample origin displayed using rarefaction curves.

Each sample is divided by treatment. Legend represents sampling months...... 181

Figure 3.9. The relative abundance of major phyla in bulk soil samples collected in May.

Each bar represents a sample within a treatment...... 182

Figure 3.10. The relative abundance of major phyla in tomato root samples. Each bar represents a sample within a treatment in a sampling month...... 183

Figure 3.11. The relative abundance of major phyla in tomato stem samples. Each bar represents a sample within a treatment in a sampling month...... 184

Figure 3.12. The relative abundance of major phyla in tomato leaf samples. Each bar represents a sample within a treatment in a sampling month...... 185

Figure 3.13. The relative abundance of major phyla in tomato flesh samples. Each bar represents a sample within a treatment in a sampling month...... 186

Figure 3.14. The relative abundance of major phyla in tomato seed samples. Each bar represents a sample within a treatment in a sampling month...... 187

Figure 3.15. The relative abundance of major phyla in tomato rhizosphere soil samples.

Each bar represents a sample within a treatment in a sampling month...... 188

Figure 3.16. The relative abundance of major phyla in chicken manure samples. Each bar represents a sample within a plot...... 189

Figure 3.17. Bacterial communities were significantly separated by sampling month and sample origin. Principal coordinate analysis of bacterial communities of rhizosphere soil, root and stem samples. Weighted normalized UniFrac distance matrix was calculated from ASV distribution...... 190

xix

Figure 3.18. The top five phyla (, Actinobacteria, Bacteroidetes, Firmicutes and Proteobacteria) were distributed among all samples with the exception of

Acidobacteria being the most clustered with rhizosphere samples. Principal coordinate analysis of bacterial communities of rhizosphere soil, root and stem samples and top five phyla. A: Distribution of bacterial communities by treatment, sample origin and sampling month. B: Distribution of top five phyla across samples. Samples that cluster close together share a greater similarity in composition...... 191

xx

Chapter 1. Literature Review

Introduction

The science that elucidates abundance, diversity, distribution, and functions of microorganisms in ecosystems is known as microbial ecology. Microorganisms are known to have the vast majority of the genetic and metabolic diversity on the planet while engaging in critical ecosystem processes (Gray & Head, 2008). Consequently, their intrinsic roles in the ecosystem allows microorganisms to evolve alongside and within multicellular forms of life, such as humans and plants. It is known that plants are influenced by the complex microbial communities they host and this shapes plant’s ecology, evolution and phenotype, a concept also known as the plant holobiont (Simon et al., 2019).

Research regarding beneficial plant-microbe interactions has been primarily focused on symbiotic associations of rhizobia, which are bacteria that form nodules in roots of legumes and fix nitrogen, and the role of arbuscular mycorrhizal fungi in nutrient uptake and plant growth (Parniske, 2008; Udvardi & Poole, 2013). In the past decades, microbial-based agricultural management efforts typically inoculate individual bacteria and fungi in crops to promote growth and development, nitrogen and phosphorous uptake, and disease resistance (Busby et al., 2017). However, these inoculants have met 1 variable success of efficacy due to various factors, such as microbial competition and survival in soil (R. de Souza et al., 2015). For example, inconsistent yield responses of single strain inoculations of Azospirillum, a plant-growth-promoting bacterial inoculant of choice, have been reported (Bashan et al., 2004). In addition, inoculation of two foreign isolates of Funneliformis mosseae successfully excluded members of the natural arbuscular mycorrhizal fungal community as root colonizers, which raises concerns about possible biodiversity losses (Pellegrino et al., 2012).

Focusing research efforts from the use of single strains to understanding and harnessing whole plant microbiomes can provide solutions for increasing plant resilience and yields in sustainable agroecosystems research. Agricultural management strategies can impact, positively or negatively, soil microbiomes, ultimately affecting the host plant’s susceptibility to disease, growth and development (Crowder et al., 2010;

Santhanam et al., 2015; Voort et al., 2016). Increased fertilization and reduced crop diversity negatively affect the abundance of bacteria in soil. Practices associated with minimizing the use of off-farm resources, such as reduced tillage regimes and crop rotations, can cause shifts in plant-associated bacteria and its surrounding agroecosystem

(Degrune et al., 2017; Schmidt et al., 2019). Depending on the practice, bacterial diversity, and in some cases abundance, in soil is promoted (Hartman et al., 2018;

Hartmann et al., 2015; Peralta et al., 2018). Sun et al. (2004) demonstrated that cattle manure application in soil increased soil bacterial diversity and had a more even distribution than soil with inorganic fertilizer applications and untreated control soil.

2

Achieving a better understanding of the impact of agricultural management strategies on the plant microbiome, specifically the beneficial functions of its endophytes, could pave way for the development of solutions to control current and emerging pathogens, as well as alleviate abiotic and biotic stresses.

Bacteria, fungi, oomycetes, algae, protozoa, nematodes, and viruses compose the plant microbiota (Hacquard et al., 2015; Jones, 2014; Kemen, 2014). Techniques from culturing isolates to metagenomic approaches, have helped shed light on uncovering the diversity and abundance of microbial species in all plant tissues and raise questions on their functions. Culture‐dependent techniques, which involves phenotypic and genomic fingerprinting analyses of bacterial isolates, can help elucidate mechanisms and physiological aspects that contribute to the microbe-host interactions within a specific ecosystem (Laforest‐Lapointe and Whitaker, 2019). With the help of next generation sequencing and statistical approaches, our understanding of the phylogeny and functional genes of bacteria has improved and has provided new understanding into the structural shifts of impacted microbial communities (Wakelin et al. 2013). Therefore, these techniques have shown that bacterial microbiomes form a vast mutualist network with their plant hosts.

The Plant Microbiome

Plants host beneficial, commensal and pathogenic microbial communities in all its compartments such as the rhizosphere, phyllosphere, spermosphere, and endosphere 3

(Turner et al. 2013). Beneficial plant-associated bacteria have demonstrated to help in disease suppression, growth promotion, and tolerance to abiotic stresses (Mueller and

Sachs, 2015). These microbes can facilitate plant growth by producing phytohormones, increasing accessibility of nutrients and protection from plant pathogens (Gaiero et al.

2013). Additionally, each plant tissue provides selective pressures that affect the assembly of microbial communities residing near, on, or inside plants. Studies on the bacterial plant microbiota have demonstrated that there are few dominant phyla, mainly

Proteobacteria, Actinobacteria, Bacteroidetes, and Firmicutes (Müller et al., 2016;

Remus‐Emsermann et al., 2014). However, differences in bacterial community composition can be observed by plant compartment. Studies on Arabidopsis thaliana and rice found significantly different communities between bulk soil, the rhizosphere, and the root compartments (Bulgarelli et al., 2012; Edwards et al., 2015; Lundberg et al., 2012).

In sugarcane, a distinction was found in bacterial communities between belowground organs (rhizosphere, endophytic root and bulk soil) and aboveground organs (stalks and leaves) (de Souza et al., 2016). A study on tomato plants found that roots have significantly enriched microbial diversity in comparison to aboveground surfaces

(Ottesen et al., 2013). The microbial communities are considered endophytic, which live inside the plant, or epiphytic, which reside on the plant surface (Bacon & White, 2016).

In the last decade, there has been an increase in research regarding bacterial endophytes, which are microbes that colonize the interior of plants without causing disease.

4

Bacterial Endophytes in Aboveground Plant Tissues

Bacterial endophytes can be found in above-ground compartments of the plant such as the stem, leaves and seeds. Endophytic communities found in each of these compartments are distinct and can reside in the vascular system, in intracellular spaces or intercellular spaces (Koskimäki et al., 2015; Thomas et al., 2007; Thomas & Sekhar,

2014; Zinniel et al., 2002). By residing in these spaces, endophytic bacteria have access to an abundance of carbohydrates, amino acids, and inorganic nutrients (Dong et al.,

1994; Elbeltagy et al., 2001). Additionally, these bacteria are more protected from environmental fluctuations, in comparison to bacteria that reside in the rhizosphere

(Rosenblueth & Martínez-Romero, 2006). However, bacteria that thrive in endophytic compartments possess physiological advantages to be able to move through the plant.

Bacterial movement can be assisted through the plant transpiration stream and having flagella (Compant et al., 2005; James et al., 2002). Secretion of cell-wall degrading enzymes like cellulases and pectinases are also helpful for bacterial movement through intercellular spaces (Compant et al., 2010; Naveed et al., 2014; Reinhold-Hurek et al.,

2006). The xylem has perforated plates that allow bacteria to pass through these large pores without using cell-wall degrading enzymes (Sapers et al., 2005).

Some of the functions described for bacterial endophytes in aboveground plant tissues include alleviating abiotic stress, defending from biotic stress and increasing accessibility of nutrients such as nitrogen, phosphorus and iron (Bacon & White, 2016; Deyett et al.,

2017). For example, four species of gfp-tagged rhizobia were observed to colonize healthy rice plant tissues, from the root to the stem and leaves, and plants produced 5 significantly higher root and shoot biomass, increased their photosynthetic rate and accumulated higher levels of indoleacetic acid and gibberellin growth-regulating phytohormones (Chi et al., 2005). And severity of angular leaf spot and naturally occurring anthracnose were significantly reduced by inoculation of Bacillus pumilus strain INR7, which was isolated from a surface-sterilized stem of a surviving cucumber plant in a field heavily infested with cucurbit wilt disease (Wei, 1996).

Bacteria in Soil and Rhizosphere of Plants

Although various physical and chemical soil factors can increase nutrient availability and plant productivity, soil microbial communities play a huge role in plant development (Bergna et al., 2018; Hartmann et al., 2015; López et al., 2018; Shaik &

Thomas, 2019). A single gram of bulk soil can contain up to 10 billion bacterial cells

(Raynaud & Nunan, 2014). On the other hand, a gram of rhizosphere soil can contain up to 100 billion bacterial cells (Kennedy & de Luna, 2005). Through altering resources, such as deposits of exudates, metabolites and induced changes to the abiotic soil environment, within the rhizosphere of an individual plant host, plant diversity is a significant driver of soil microbial community structure, richness and interactions

(Bardgett & Wardle, 2010; Schlatter et al., 2015; Wardle et al., 2004). Land use, including agricultural management strategies, soil resource stoichiometry, climate, soil spatial heterogeneity and soil pH are other factors that can impact bacterial diversity and composition in soil (Delgado‐Baquerizo et al., 2017; Ding et al., 2013; Wang et al.,

2016). 6

The rhizosphere is the area where plant roots deposit exudates, mucilage, and sloughed cells, that influence the surrounding soil and the large array and diversity of organisms within (Lakshmanan et al., 2014). Microorganisms in the rhizosphere that have been studied for their beneficial effects on plant growth and health include nitrogen- fixing bacteria, mycorrhizal fungi, plant growth-promoting rhizobacteria, biocontrol microorganisms, mycoparasitic fungi, and protozoa (Zinniel et al., 2002). The rhizosphere hosts a plethora of bacteria with plant growth promoting and disease suppressive abilities (Adesemoye et al., 2009; Srour et al., 2017; Weller et al., 2002).

Plants are known to manipulate the bulk soil and rhizosphere microbiome to their advantage, by recruiting beneficial microorganisms. Mechanisms plants employ to recruit beneficial bacteria include the production of plant root exudates and volatile organic compounds. Various studies have observed how these exudates attract specific bacteria with functions ranging from plant growth promotion to production of pathogen-inhibitory compounds (Rudrappa et al., 2008; Schulz-Bohm et al., 2018). On the other hand, these compounds can also signal plant pathogens (Walker et al., 2003). In a study by Rudrappa et al. (2008), the authors provide biochemical evidence that malic acid secreted from roots of Arabidopsis thaliana selectively signals and recruits the beneficial rhizobacterium Bacillus subtilis FB17 in a dose-dependent manner after with foliar pathogen syringae pv. tomato. Different strains of B. subtilis have also been reported to produce cytokinins and increases the phytohormone content in the plants (Arkhipova et al., 2005).

7

Transmission and Colonization of Plant-Associated Bacterial Communities

To understand the role endophytes play in plants and its ecosystem, it is important to understand how endophytic communities assemble. Although plant compartments are studied separately, it is important to remember that each of these systems can share bacterial communities (Bai et al., 2015). Endophytes are transmitted horizontally, by entering the plant from an environmental origin, or vertically, from generation to generation via seed (Frank et al., 2017). Sources for horizontal transmission of bacterial endophytes include the air, rain, soil, pollinators or other insects. The aerial surface of plants, which include stem, leaves, flowers and fruits, can have bacteria deposited from bioaerosols, which are particles released by wind from terrestrial and marine environments into the atmosphere and include viable and metabolically active bacteria

(Fröhlich-Nowoisky et al., 2016; Hill et al., 2007). The surfaces of floral organs are visited by pollinators, herbivores and predators in search of prey (Wardhaugh et al.,

2012). In turn, these visits influence floral microbiomes in areas such as nectar, petals, pollen and fruit (Aizenberg-Gershtein et al., 2013; Fürnkranz et al., 2012; Glassner et al.,

2015; Madmony et al., 2005; McFrederick et al., 2017; Ushio et al., 2015).

Although, bacteria colonizing the surface of leaves do not penetrate the leaf cuticle, bacteria can enter through openings in the plant epidermis such as stomata, lenticels and hydathodes (Hugouvieux et al., 1998; Melotto et al., 2008; Scott et al.,

1996). The stomata is known to recognize and exclude specific bacteria, as well as allow non-pathogenic bacteria to enter (Baldotto et al., 2011). Also known as guard cells, the 8 stomata can sense microbe-associated molecular patterns (MAMPs) and close in response

(Arnaud & Hwang, 2015). Pathogens are known to possess mechanisms to open the stomatal pore through virulence factors, including phytotoxins that prevent MAMP- triggered closure and effectors that prevent closure or induce opening of the stomata

(Hurley et al., 2014; Melotto et al., 2006). Diazotrophic species of bacterial endophytes have similarly shown to alter stomatal opening. A study demonstrated the effect of abscisic acid (ABA), key hormone of stomatal control and its diurnal fluctuation, production by diazotrophs, such as Rhizobium and Burkholderia, in assisting in water use efficiency of rice (Rho et al., 2018; Tallman, 2004).

Some endophytes spread systemically to various areas of the plant like stem, leaves, and fruits via the xylem (Compant et al., 2010; James et al., 2002). Bacterial endophytes can move between vegetative parts of the plant, including the developing seed, through vascular connections (Agarwal & Sinclair, 1996). These bacteria can also move via the shoot apical meristem, which consists of cells that give rise to all the post- embryonic aerial organs, including reproductive organs (Clark, 1997). For example, leaf- nodulating nitrogen-fixing Burkholderia symbionts reside in the vegetative shoot tip and colonize each new leaf in Rubiaceae species (Sinnesael et al., 2018). Burkholderia symbionts are then transferred into the floral shoot tip, into the developing ovule and eventually the embryo where they become enclosed in the shoot tip of the seedling

(Miller, 1990). The movement of the endophyte B. phytofirmans strain PsJN from the rhizosphere to the grapevine’s reproductive organs following soil inoculation was studied

9

(Compant et al., 2008). This study helped elucidate an example of transmission of a plant growth promoting bacteria, originating from the soil, via xylem colonization.

The soil contains a microbial seed bank which is a plethora of active microorganisms, as well as dormant ones (Lennon & Jones, 2011). The rhizosphere creates a selective environment that supports higher bacterial abundance and activity compared to other plant compartments and bulk soil, while at the same time lowering bacterial diversity (Kowalchuk et al., 2002; Smalla et al., 2001). In turn, the rhizosphere recruits distinct microbial communities that influence the structure and assembly of communities transmitted to the inside of the root (Peiffer et al., 2013). Bacteria, including plant pathogens, can persist and overwinter in plant residues (Togashi et al., 2001).

Bacteria enter the roots through cracks, root tips and wounds, where lateral roots emerge from (Hurek et al., 1994). For this, bacteria engage in mechanisms to approach and colonize roots and compete with other microorganisms, such as motility, chemotaxis and quorum sensing (Podile et al., 2014; Scharf et al., 2016). Additionally, bacteria can secrete cell wall-degrading enzymes, such as endoglucanase and endopolygalacturonase, to assist in penetration into the root endosphere (Compant et al., 2005). These colonization routes are similar between bacterial endophytes and bacterial plant pathogens. The bacterial plant pathogen Ralstonia solanacearum causes bacterial wilt, a soil-borne vascular disease that affects many economically important crops, such as tomato. R. solanacearum can survive and overwinter in plant root debris, or roots of weeds (Pradhanang et al., 2000). Similar to a bacterial endophyte, R. solanacearum uses

10 plant cell wall-degrading enzymes to colonize and enter the plant host (Digonnet et al.,

2012). Virulent strains of R. solanacearum compared with mutant strains lacking production of these enzymes have shown reduced virulence in tomato plants (Huang &

Allen, 2000; Saile et al., 1997). Other than roots, bacteria are known to colonize germinating seeds in the soil. The zone surrounding a germinating seed is known as the spermosphere (Schiltz et al., 2015). The zone is composed of carbon exudates in the form of sugars, proteins, and fatty acids that serve as energy sources for bacteria, eventually shaping the bacterial composition near the seed (Roberts et al., 2009).

Sources for vertical transmission of bacterial endophytes include the seed and pollen. One of the least researched plant tissues is the seed and its indigenous microbial communities. Bacterial seed endophytes can be selected by the soil microbial community or the seed genotype and serve as a vehicle for vertical transmission through generations

(Nelson, 2004). After seed germination, genetic and physiological changes in the plant contribute to significant differences in bacterial endophyte colonization that can vary between cultivars (Adams and Kloepper, 2002). Certain seed endophytes have shown to be beneficial to the host, through plant growth promotion, antibiotic production and promote resilience to stress (Shahzad et al. 2018). Considering sources for horizontal transmission, it has been difficult to prove vertical transfer of obligate bacterial endophytes. It has also been difficult to obtain culturable bacterial endophytes from the seed. A study found that surface-sterilized seeds hosted only one PDA-culturable bacterial or fungal endophyte per seed, or none (Newcombe et al., 2018). The authors in

11 this study could explain this occurrence with the Primary Symbiont Hypothesis. This hypothesis states that plants host primary, sole, or dominant, endophytic symbionts in individual seeds and its presence or absence and identity of the primary symbiont affects survival of the host during seed dispersal, germination, emergence and young growth

(Mundt & Hinkle, 1976). Although culture-independent techniques reveal several operational taxonomic units per seed, it is thought that a primary symbiont could be as dominant as to have a larger effect on fitness of seedlings (Truyens et al., 2015).

Role of Bacterial Endophytes

Endophytic bacteria inhabit plants without causing harm. They are known to mediate, enhance or alter plant functional traits that influence the plant’s interactions with competitors, mutualists, herbivores, and pathogens and can play a role in niche range, abundance, growth and development (Friesen et al., 2011). Bacterial endophytes can be efficient for phytoremediation (Barac et al., 2004). Plant endophytes have also been found to be closely related to human pathogens or are human pathogens. For example, plant growth promoting traits have been identified in Bacillus cereus, a plant endophyte also known as a human pathogen, associated to food poisoning (Bottone, 2010; Vendan et al., 2010).

A study compared the genomes of 40 well-described plant-associated bacteria, including endophytes, to understand their potential functional and mechanistic interactions (Hardoim et al., 2015) . The 40 bacterial strains studied included strict 12 endophytes, symbionts, plant pathogens with endophytic lifestyle, rhizosphere and soil bacteria. They found that motility and chemotaxis and the generation of detoxification and stress-related enzymes were important properties driving colonization of bacteria in plants. The plant endosphere can be inhospitable for aerobic microorganisms due to it being a microaerophilic environment (Das & Roychoudhury, 2014). Therefore, endophytes need to be able to generate enzymes with detoxification capacities to survive in plants. Genes associated to type IV conjugal DNA-protein transfer secretion systems, which are involved with host colonization and conjugation of DNA, adhesion to the host via twitching motility and type I pilus assembly, which assists in attaching to host cells, were detected among endophytes. Furthermore, genes involved in plant growth promotion, such as nitrogenase (nifH), involved in fixation of atmospheric N2, and 1- aminocyclopropane-1-carboxylase (ACC) deaminase (acdS), involved in plant stress alleviation, were more common in endophytes. However, functions associated to signal transduction, bacterial cell communication, nutrient transport and transcriptional regulators varied between functional groups depending on what suited their survival in different environmental niches and conditions. Another study compared 3,837 bacterial genomes to identify genes that contribute to bacterial adaptation to plants (Levy et al.,

2018). The authors found numerous plant-associated functions, such as those that facilitate plant colonization, consistent across phylogenetically diverse bacterial taxa, and shared with plant-associated eukaryotes.

13

Plant-growth promoting activities have been mostly studied on bacteria associated with the rhizosphere (Berendsen et al., 2012; Kwak et al., 2018; Lugtenberg & Kamilova,

2009; Zhao et al., 2017). However, there are examples of endophytes shown to promote plant growth (Miliute et al., 2015). In the case of legumes, they can host rhizobia, which are nitrogen-fixing bacteria, in nodules in their roots, with in turn help plants in nitrogen limited environments (Oldroyd, 2013). B. phytofirmans PsJN is a well characterized plant growth promoting endophyte. Mutants of this bacteria lacking ACC deaminase activity ceased to promote canola seedling root elongation (Y. Sun et al., 2009). Endophytic bacteria can also increase host tolerance to abiotic stresses (Bresson et al., 2013;

Coleman-Derr & Tringe, 2014; Lucas et al., 2014). YsS6 and

Pseudomonas migulae 8R6, endophytes with ACC deaminase activity, promoted tomato plant growth in high salinity (Ali et al., 2014). Endophytic Azospirillum spp. alleviated water stress tolerance in maize plants by producing ABA and promoted plant growth through the production of indole acetic acid and gibberellins (Cohen et al., 2009).

Through various mechanisms, endophytic bacteria can decrease or prevent the destructive effects of bacterial and fungal plant pathogens (Eljounaidi et al., 2016).

Endophytic Pseudomonas fluorescens was shown to produce volatile organic compounds, antibiotics and lytic exoenzymes that inhibit rolfsii, a soil‐borne fungal pathogen of Atractylodes lancea (Zhou et al., 2014). Seed endophyte Bacillus subtilis SG_JW.03 suppressed fungal pathogen Fusarium moniliforme growth on Arabidopsis thaliana seedlings through the production of lipopeptides and showed the induction of

14 pathogenesis-related genes, including ones that relate to plant defense against fungal pathogens (Gond et al., 2015). Pavlo et al. (2011) demonstrated that Pseudomonas sp. increased growth and resistance of potato plant shoots towards soft rot disease caused by

Pectobacterium atrosepticum.

Summary and Conclusions

Gaps in research should be considered when understanding the effects of bacterial microbiome on host’s performance. A specific group of endophytes that tends to be overlooked are the commensals, which are the most common endophytes with unknown or yet unknown functions in plants (Berlec, 2012). Increasing knowledge of the functionality of this group can shift our understanding on the role of endophytic communities and how these might me impacting the plant host. Studies on bacterial communities in agroecosystems have been mainly focused on the rhizosphere. This is due to the fact that intensively managed soils can affect soil health and microbial abundance and diversity, ultimately decreasing disease suppression (Chaparro et al., 2012; Havlin et al., 1990; Mendes et al., 2011; Voort et al., 2016). For example, high use of pesticides and fungicides used to target plant pathogens can reduce soil microbial community diversity and evenness (Liu et al., 2007). However, there are not as many studies on the impact this might have on above-ground endophytes. Moreover, there is a lack of research on model non-pathogenic bacterial endophytes and how these factors impact their colonization, horizontal and vertical transmission, and function in plants. Most of

15 the research used to understand endophytic lifestyle is obtained from plant pathogenic bacteria.

When studying drivers in endophyte colonization, many experiments grow host plants under controlled conditions (Cordovez et al., 2018; Hardoim et al., 2012). While this is necessary, there is a lack of studies assessing variable field conditions that might impact endophytic community composition and function. This premise can apply to studies involving agroecosystems. Additionally, there is a lack of studies regarding the impact of management strategies associated to diversified farming systems, such as those that employ crop-livestock rotation, on the plant microbiome. The development of high yielding varieties of crops, increased use of fertilizers and other management strategies associated to the “green revolution” have resulted in lack of sustainability, loss of biodiversity and pollution (Lemaire et al. 2005). It has been found that crop-livestock systems can increase soil quality and fertility, reduce cost of herbicide use and improve sustainability, especially for farmers in poorer areas of the world (de Faccio Carvalho et al. 2010; Devendra and Thomas, 2002). Studies involving the impact of this type of management strategies on microbial communities are not common. One study showed an increase in microbial biomass and enzyme activities in a crop-livestock system compared to a monoculture (Acosta-Martínez et al. 2010). Although crop-livestock systems have demonstrated their efficiency and improved sustainability, the impact of this strategy on the plant microbiome is not completely known. It is important to address these gaps in knowledge to be able to understand this complex ecological network of bacterial

16 endophytes in plants and how the information gained can be transformed into sustainable solutions for agroecosystems.

For this thesis, a characterization of the endophytic bacterial microbiome of the tomato plant and its surrounding rhizosphere soil was done by using culture-dependent and culture independent approaches. The aim of this research project was to explore what influences bacterial diversity in soil and tomato plant tissues at different stages of its life cycle, using an integrated crop and livestock systems with chickens as a model.

The two objectives for this thesis are as follow:

1. Characterize the cultured bacterial endophytic microbiome of tomato

seedlings grown in a soil diversity gradient using chicken-grazed soils.

2. Characterize the endophytic bacterial microbiome of tomato plants

grown until harvest in a chicken grazed plot.

We hypothesize that bacterial endophyte diversity will increase in chicken-grazed soils compared to soils with pasture, and that endophyte diversity will decrease (and be more similar to the seed source) as soil biodiversity decreases. This project will help understand the potential of integrating crop-livestock strategies from the perspective of tomato endophyte diversity and beneficial effects in tomato growth. Elucidating more about the plant microbiome can pave way in the development of sustainable agricultural practices and eventually, employ these systems, close information gaps, increase energy

17 efficiency and decrease negative environmental impact to further enhance food production.

References

Acosta-Martínez, V., Bell, C. W., Morris, B. E. L., Zak, J., & Allen, V. G. (2010). Long-term soil microbial community and enzyme activity responses to an integrated cropping-livestock system in a semi-arid region☆. Agriculture, Ecosystems & Environment, 137(3–4), 231–240. https://doi.org/10.1016/j.agee.2010.02.008

Adams, P. D., & Kloepper, J. W. (2002). Effect of host genotype on indigenous bacterial endophytes of cotton (Gossypium hirsutum L.). Plant and Soil, 240(1), 181-189.

Adesemoye, A. O., Torbert, H. A., & Kloepper, J. W. (2009). Plant Growth-Promoting Rhizobacteria Allow Reduced Application Rates of Chemical Fertilizers. Microbial Ecology, 58(4), 921–929. https://doi.org/10.1007/s00248-009-9531-y

Agarwal, V. K., & Sinclair, J. B. (1996). Principles of Seed Pathology, Second Edition (2 edition). Lewis Publishers.

Aizenberg-Gershtein, Y., Izhaki, I., & Halpern, M. (2013). Do Honeybees Shape the Bacterial Community Composition in Floral Nectar? PLOS ONE, 8(7), e67556. https://doi.org/10.1371/journal.pone.0067556

Ali, S., Charles, T. C., & Glick, B. R. (2014). Amelioration of high salinity stress damage by plant growth-promoting bacterial endophytes that contain ACC deaminase. Plant Physiology and Biochemistry, 80, 160–167. https://doi.org/10.1016/j.plaphy.2014.04.003

Arkhipova, T. N., Veselov, S. U., Melentiev, A. I., Martynenko, E. V., & Kudoyarova, G. R. (2005). Ability of bacterium Bacillus subtilis to produce cytokinins and to influence the growth and endogenous hormone content of lettuce plants. Plant and Soil, 272(1), 201–209. https://doi.org/10.1007/s11104-004-5047-x

Arnaud, D., & Hwang, I. (2015). A sophisticated network of signaling pathways regulates stomatal defenses to bacterial pathogens. Molecular Plant, 8(4), 566– 581. https://doi.org/10.1016/j.molp.2014.10.012

18

Bacon, C. W., & White, J. F. (2016). Functions, mechanisms and regulation of endophytic and epiphytic microbial communities of plants. Symbiosis, 68(1–3), 87–98. https://doi.org/10.1007/s13199-015-0350-2

Bai, Y., Müller, D. B., Srinivas, G., Garrido-Oter, R., Potthoff, E., Rott, M., Dombrowski, N., Münch, P. C., Spaepen, S., Remus-Emsermann, M., Hüttel, B., McHardy, A. C., Vorholt, J. A., & Schulze-Lefert, P. (2015). Functional overlap of the Arabidopsis leaf and root microbiota. Nature, 528(7582), 364–369. https://doi.org/10.1038/nature16192

Baldotto, L. E. B., Olivares, F. L., & Bressan-Smith, R. (2011). Structural interaction between GFP-labeled diazotrophic endophytic bacterium Herbaspirillum seropedicae RAM10 and pineapple plantlets “Vitória.” Brazilian Journal of Microbiology, 42(1), 114–125. https://doi.org/10.1590/S1517- 83822011000100015

Barac, T., Taghavi, S., Borremans, B., Provoost, A., Oeyen, L., Colpaert, J. V., Vangronsveld, J., & van der Lelie, D. (2004). Engineered endophytic bacteria improve phytoremediation of water-soluble, volatile, organic pollutants. Nature Biotechnology, 22(5), 583–588. https://doi.org/10.1038/nbt960

Bardgett, R. D., & Wardle, D. A. (2010). Aboveground-belowground linkages: Biotic interactions, ecosystem processes, and global change. Oxford University Press.

Bashan, Y., Holguin, G., & de-Bashan, L. E. (2004). Azospirillum-plant relationships: Physiological, molecular, agricultural, and environmental advances (1997- 2003). Canadian Journal of Microbiology; Ottawa, 50(8), 521–577.

Berendsen, R. L., Pieterse, C. M. J., & Bakker, P. A. H. M. (2012). The rhizosphere microbiome and plant health. Trends in Plant Science, 17(8), 478–486. https://doi.org/10.1016/j.tplants.2012.04.001

Bergna, A., Cernava, T., Rändler, M., Grosch, R., Zachow, C., & Berg, G. (2018). Tomato Seeds Preferably Transmit Plant Beneficial Endophytes. Phytobiomes Journal, 2(4), 183–193. https://doi.org/10.1094/PBIOMES-06-18-0029-R

Berlec, A. (2012). Novel techniques and findings in the study of plant microbiota: Search for plant probiotics. Plant Science, 193–194, 96–102. https://doi.org/10.1016/j.plantsci.2012.05.010

Bottone, E. J. (2010). Bacillus cereus, a Volatile Human Pathogen. Clinical Microbiology Reviews, 23(2), 382–398. https://doi.org/10.1128/CMR.00073-09

19

Bresson, J., Varoquaux, F., Bontpart, T., Touraine, B., & Vile, D. (2013). The PGPR strain Phyllobacterium brassicacearum STM196 induces a reproductive delay and physiological changes that result in improved drought tolerance in Arabidopsis. New Phytologist, 200(2), 558–569. https://doi.org/10.1111/nph.12383

Bulgarelli, D., Rott, M., Schlaeppi, K., Ver Loren van Themaat, E., Ahmadinejad, N., Assenza, F., Rauf, P., Huettel, B., Reinhardt, R., Schmelzer, E., Peplies, J., Gloeckner, F. O., Amann, R., Eickhorst, T., & Schulze-Lefert, P. (2012). Revealing structure and assembly cues for Arabidopsis root-inhabiting bacterial microbiota. Nature, 488(7409), 91–95. https://doi.org/10.1038/nature11336

Busby, P. E., Soman, C., Wagner, M. R., Friesen, M. L., Kremer, J., Bennett, A., Morsy, M., Eisen, J. A., Leach, J. E., & Dangl, J. L. (2017). Research priorities for harnessing plant microbiomes in sustainable agriculture. PLOS Biology, 15(3), e2001793. https://doi.org/10.1371/journal.pbio.2001793

Chaparro, J. M., Sheflin, A. M., Manter, D. K., & Vivanco, J. M. (2012). Manipulating the soil microbiome to increase soil health and plant fertility. Biology and Fertility of Soils, 48(5), 489–499. https://doi.org/10.1007/s00374-012-0691-4

Chi, F., Shen, S.-H., Cheng, H.-P., Jing, Y.-X., Yanni, Y. G., & Dazzo, F. B. (2005). Ascending Migration of Endophytic Rhizobia, from Roots to Leaves, inside Rice Plants and Assessment of Benefits to Rice Growth Physiology. Applied and Environmental Microbiology, 71(11), 7271–7278. https://doi.org/10.1128/AEM.71.11.7271-7278.2005

Clark, S. E. (1997). Organ Formation at the Vegetative Shoot Meristem. The Plant Cell, 9(7), 1067–1076. https://doi.org/10.1105/tpc.9.7.1067

Cohen, A. C., Travaglia, C. N., Bottini, R., & Piccoli, P. N. (2009). Participation of abscisic acid and gibberellins produced by endophytic Azospirillum in the alleviation of drought effects in maize. Botany, 87(5), 455–462. https://doi.org/10.1139/B09-023

Coleman-Derr, D., & Tringe, S. G. (2014). Building the crops of tomorrow: Advantages of symbiont-based approaches to improving abiotic stress tolerance. Frontiers in Microbiology, 5. https://doi.org/10.3389/fmicb.2014.00283

Compant, S., Clément, C., & Sessitsch, A. (2010). Plant growth-promoting bacteria in the rhizo- and endosphere of plants: Their role, colonization, mechanisms involved and prospects for utilization. Soil Biology and Biochemistry, 42(5), 669–678. https://doi.org/10.1016/j.soilbio.2009.11.024

20

Compant, S., Kaplan, H., Sessitsch, A., Nowak, J., Ait Barka, E., & Clément, C. (2008). Endophytic colonization of Vitis vinifera L. by Burkholderia phytofirmans strain PsJN: From the rhizosphere to inflorescence tissues. FEMS Microbiology Ecology, 63(1), 84–93. https://doi.org/10.1111/j.1574-6941.2007.00410.x

Compant, S., Reiter, B., Sessitsch, A., Nowak, J., Clément, C., & Barka, E. A. (2005). Endophytic Colonization of Vitis vinifera L. by Plant Growth-Promoting Bacterium Burkholderia sp. Strain PsJN. Applied and Environmental Microbiology, 71(4), 1685–1693. https://doi.org/10.1128/AEM.71.4.1685- 1693.2005

Cordovez, V., Schop, S., Hordijk, K., Boulois, H. D. de, Coppens, F., Hanssen, I., Raaijmakers, J. M., & Carrión, V. J. (2018). Priming of Plant Growth Promotion by Volatiles of Root-Associated Microbacterium spp. Applied and Environmental Microbiology, 84(22). https://doi.org/10.1128/AEM.01865-18

Crowder, D. W., Northfield, T. D., Strand, M. R., & Snyder, W. E. (2010). Organic agriculture promotes evenness and natural pest control. Nature, 466(7302), 109– 112. https://doi.org/10.1038/nature09183

Das, K., & Roychoudhury, A. (2014). Reactive oxygen species (ROS) and response of antioxidants as ROS-scavengers during environmental stress in plants. Frontiers in Environmental Science, 2. https://doi.org/10.3389/fenvs.2014.00053 de Faccio Carvalho, P. C., Anghinoni, I., de Moraes, A., de Souza, E. D., Sulc, R. M., Lang, C. R., Flores, J. P. C., Terra Lopes, M. L., da Silva, J. L. S., Conte, O., de Lima Wesp, C., Levien, R., Fontaneli, R. S., & Bayer, C. (2010). Managing grazing animals to achieve nutrient cycling and soil improvement in no-till integrated systems. Nutrient Cycling in Agroecosystems, 88(2), 259–273. https://doi.org/10.1007/s10705-010-9360-x de Souza, R. S. C., Okura, V. K., Armanhi, J. S. L., Jorrín, B., Lozano, N., da Silva, M. J., González-Guerrero, M., de Araújo, L. M., Verza, N. C., Bagheri, H. C., Imperial, J., & Arruda, P. (2016). Unlocking the bacterial and fungal communities assemblages of sugarcane microbiome. Scientific Reports, 6(1), 28774. https://doi.org/10.1038/srep28774

Degrune, F., Theodorakopoulos, N., Colinet, G., Hiel, M.-P., Bodson, B., Taminiau, B., Daube, G., Vandenbol, M., & Hartmann, M. (2017). Temporal Dynamics of Soil Microbial Communities below the Seedbed under Two Contrasting Tillage Regimes. Frontiers in Microbiology, 8. https://doi.org/10.3389/fmicb.2017.01127

21

Delgado‐Baquerizo, M., Reich, P. B., Khachane, A. N., Campbell, C. D., Thomas, N., Freitag, T. E., Al‐Soud, W. A., Sørensen, S., Bardgett, R. D., & Singh, B. K. (2017). It is elemental: Soil nutrient stoichiometry drives bacterial diversity. Environmental Microbiology, 19(3), 1176–1188. https://doi.org/10.1111/1462- 2920.13642

Devendra, C., & Thomas, D. (2002). Crop–animal interactions in mixed farming systems in Asia. Agricultural Systems, 71(1–2), 27–40. https://doi.org/10.1016/S0308-521X(01)00034-8

Deyett, E., Roper, M. C., Ruegger, P., Yang, J.-I., Borneman, J., & Rolshausen, P. E. (2017). Microbial Landscape of the Grapevine Endosphere in the Context of Pierce’s Disease. Phytobiomes Journal, 1(3), 138–149. https://doi.org/10.1094/PBIOMES-08-17-0033-R

Digonnet, C., Martinez, Y., Denancé, N., Chasseray, M., Dabos, P., Ranocha, P., Marco, Y., Jauneau, A., & Goffner, D. (2012). Deciphering the route of Ralstonia solanacearum colonization in Arabidopsis thaliana roots during a compatible interaction: Focus at the plant cell wall. Planta: An International Journal of Plant Biology; Heidelberg, 236(5), 1419–1431. http://dx.doi.org.proxy.lib.ohio-state.edu/10.1007/s00425-012-1694-y Ding, G.-C., Piceno, Y. M., Heuer, H., Weinert, N., Dohrmann, A. B., Carrillo, A., Andersen, G. L., Castellanos, T., Tebbe, C. C., & Smalla, K. (2013). Changes of Soil Bacterial Diversity as a Consequence of Agricultural Land Use in a Semi- Arid Ecosystem. PLOS ONE, 8(3), e59497. https://doi.org/10.1371/journal.pone.0059497

Dong, Z., Canny, M. J., McCully, M. E., Roboredo, M. R., Cabadilla, C. F., Ortega, E., & Rodes, R. (1994). A Nitrogen-Fixing Endophyte of Sugarcane Stems (A New Role for the Apoplast). Plant Physiology, 105(4), 1139–1147. https://doi.org/10.1104/pp.105.4.1139

Edwards, J., Johnson, C., Santos-Medellín, C., Lurie, E., Podishetty, N. K., Bhatnagar, S., Eisen, J. A., & Sundaresan, V. (2015). Structure, variation, and assembly of the root-associated microbiomes of rice. Proceedings of the National Academy of Sciences of the United States of America, 112(8), E911-920. https://doi.org/10.1073/pnas.1414592112

Elbeltagy, A., Nishioka, K., Sato, T., Suzuki, H., Ye, B., Hamada, T., Isawa, T., Mitsui, H., & Minamisawa, K. (2001). Endophytic Colonization and In Planta Nitrogen Fixation by a Herbaspirillum sp. Isolated from Wild Rice Species. Applied and Environmental Microbiology, 67(11), 5285–5293. https://doi.org/10.1128/AEM.67.11.5285-5293.2001

22

Eljounaidi, K., Lee, S. K., & Bae, H. (2016). Bacterial endophytes as potential biocontrol agents of vascular wilt diseases – Review and future prospects. Biological Control, 103, 62–68. https://doi.org/10.1016/j.biocontrol.2016.07.013

Frank, A., Saldierna Guzmán, J., & Shay, J. (2017). Transmission of Bacterial Endophytes. Microorganisms, 5(4), 70. https://doi.org/10.3390/microorganisms5040070

Friesen, M. L., Porter, S. S., Stark, S. C., von Wettberg, E. J., Sachs, J. L., & Martinez- Romero, E. (2011). Microbially Mediated Plant Functional Traits. Annual Review of Ecology, Evolution, and Systematics, 42(1), 23–46. https://doi.org/10.1146/annurev-ecolsys-102710-145039

Fröhlich-Nowoisky, J., Kampf, C. J., Weber, B., Huffman, J. A., Pöhlker, C., Andreae, M. O., Lang-Yona, N., Burrows, S. M., Gunthe, S. S., Elbert, W., Su, H., Hoor, P., Thines, E., Hoffmann, T., Després, V. R., & Pöschl, U. (2016). Bioaerosols in the Earth system: Climate, health, and ecosystem interactions. Atmospheric Research, 182, 346–376. https://doi.org/10.1016/j.atmosres.2016.07.018

Fürnkranz, M., Lukesch, B., Müller, H., Huss, H., Grube, M., & Berg, G. (2012). Microbial Diversity Inside Pumpkins: Microhabitat-Specific Communities Display a High Antagonistic Potential Against Phytopathogens. Microbial Ecology, 63(2), 418–428. https://doi.org/10.1007/s00248-011-9942-4

Gaiero, J. R., McCall, C. A., Thompson, K. A., Day, N. J., Best, A. S., & Dunfield, K. E. (2013). Inside the root microbiome: Bacterial root endophytes and plant growth promotion. American Journal of Botany, 100(9), 1738–1750. https://doi.org/10.3732/ajb.1200572

Glassner, H., Zchori-Fein, E., Compant, S., Sessitsch, A., Katzir, N., Portnoy, V., & Yaron, S. (2015). Characterization of endophytic bacteria from cucurbit fruits with potential benefits to agriculture in melons (Cucumis melo L.). FEMS Microbiology Ecology, 91(7). https://doi.org/10.1093/femsec/fiv074

Gond, S. K., Bergen, M. S., Torres, M. S., & White Jr, J. F. (2015). Endophytic Bacillus spp. Produce antifungal lipopeptides and induce host defence gene expression in maize. Microbiological Research, 172, 79–87. https://doi.org/10.1016/j.micres.2014.11.004 Gray, N. D., & Head, I. M. (2008). Microbial Ecology. In Encyclopedia of Ecology (pp. 2357–2368). Academic Press. https://doi.org/10.1016/B978-008045405- 4.00519-X

Hacquard, S., Garrido-Oter, R., González, A., Spaepen, S., Ackermann, G., Lebeis, S., McHardy, A. C., Dangl, J. L., Knight, R., Ley, R., & Schulze-Lefert, P. (2015). 23

Microbiota and Host Nutrition across Plant and Animal Kingdoms. Cell Host & Microbe, 17(5), 603–616. https://doi.org/10.1016/j.chom.2015.04.009

Hardoim, P. R., Hardoim, C. C. P., Overbeek, L. S. van, & Elsas, J. D. van. (2012). Dynamics of Seed-Borne Rice Endophytes on Early Plant Growth Stages. PLOS ONE, 7(2), e30438. https://doi.org/10.1371/journal.pone.0030438

Hardoim, P. R., van Overbeek, L. S., Berg, G., Pirttilä, A. M., Compant, S., Campisano, A., Döring, M., & Sessitsch, A. (2015). The Hidden World within Plants: Ecological and Evolutionary Considerations for Defining Functioning of Microbial Endophytes. Microbiology and Molecular Biology Reviews, 79(3), 293–320. https://doi.org/10.1128/MMBR.00050-14

Hartman, K., van der Heijden, M. G. A., Wittwer, R. A., Banerjee, S., Walser, J.-C., & Schlaeppi, K. (2018). Cropping practices manipulate abundance patterns of root and soil microbiome members paving the way to smart farming. Microbiome, 6(1), 14. https://doi.org/10.1186/s40168-017-0389-9

Hartmann, M., Frey, B., Mayer, J., Mäder, P., & Widmer, F. (2015). Distinct soil microbial diversity under long-term organic and conventional farming. The ISME Journal, 9(5), 1177–1194. https://doi.org/10.1038/ismej.2014.210

Havlin, J. L., Kissel, D. E., Maddux, L. D., Claassen, M. M., & Long, J. H. (1990). Crop Rotation and Tillage Effects on Soil Organic Carbon and Nitrogen. Soil Science Society of America Journal, 54(2), 448–452. https://doi.org/10.2136/sssaj1990.03615995005400020026x

Hill, K. A., Shepson, P. B., Galbavy, E. S., Anastasio, C., Kourtev, P. S., Konopka, A., & Stirm, B. H. (2007). Processing of atmospheric nitrogen by clouds above a forest environment. Journal of Geophysical Research: Atmospheres, 112(D11). https://doi.org/10.1029/2006JD008002

Huang, Q., & Allen, C. (2000). Polygalacturonases are required for rapid colonization and full virulence of Ralstonia solanacearum on tomato plants. Physiological and Molecular Plant Pathology, 57(2), 77–83. https://doi.org/10.1006/pmpp.2000.0283

Hugouvieux, V., Barber, C. E., & Daniels, M. J. (1998). Entry of Xanthomonas campestris pv. campestris into Hydathodes of Arabidopsis thaliana Leaves: A System for Studying Early Infection Events in Bacterial Pathogenesis. Molecular Plant-Microbe Interactions®, 11(6), 537–543. https://doi.org/10.1094/MPMI.1998.11.6.537

24

Hurek, T., Reinhold-Hurek, B., Van Montagu, M., & Kellenberger, E. (1994). Root colonization and systemic spreading of Azoarcus sp. Strain BH72 in grasses. Journal of Bacteriology, 176(7), 1913–1923.

Hurley, B., Lee, D., Mott, A., Wilton, M., Liu, J., Liu, Y. C., Angers, S., Coaker, G., Guttman, D. S., & Desveaux, D. (2014). The Pseudomonas syringae type III effector HopF2 suppresses Arabidopsis stomatal immunity. PloS One, 9(12), e114921. https://doi.org/10.1371/journal.pone.0114921

James, E. K., Gyaneshwar, P., Mathan, N., Barraquio, W. L., Reddy, P. M., Iannetta, P. P. M., Olivares, F. L., & Ladha, J. K. (2002). Infection and Colonization of Rice Seedlings by the Plant Growth-Promoting Bacterium Herbaspirillum seropedicae Z67. Molecular Plant-Microbe Interactions®, 15(9), 894–906. https://doi.org/10.1094/MPMI.2002.15.9.894

Jones, R. a. C. (2014). Plant virus ecology and epidemiology: Historical perspectives, recent progress and future prospects. Annals of Applied Biology, 164(3), 320– 347. https://doi.org/10.1111/aab.12123

Kemen, E. (2014). Microbe–microbe interactions determine oomycete and fungal host colonization. Current Opinion in Plant Biology, 20, 75–81. https://doi.org/10.1016/j.pbi.2014.04.005

Kennedy, A. C., & de Luna, L. Z. (2005). Rhizosphere. In D. Hillel (Ed.), Encyclopedia of Soils in the Environment (pp. 399–406). Elsevier. https://doi.org/10.1016/B0- 12-348530-4/00163-6

Koskimäki, J. J., Pirttilä, A. M., Ihantola, E.-L., Halonen, O., & Frank, A. C. (2015). The intracellular Scots pine shoot symbiont Methylobacterium extorquens DSM13060 aggregates around the host nucleus and encodes eukaryote-like proteins. MBio, 6(2). https://doi.org/10.1128/mBio.00039-15

Kowalchuk, George. A., Buma, D. S., de Boer, W., Klinkhamer, P. G. L., & van Veen, J. A. (2002). Effects of above-ground plant species composition and diversity on the diversity of soil-borne microorganisms. Antonie van Leeuwenhoek, 81(1), 509. https://doi.org/10.1023/A:1020565523615

Kwak, M.-J., Kong, H. G., Choi, K., Kwon, S.-K., Song, J. Y., Lee, J., Lee, P. A., Choi, S. Y., Seo, M., Lee, H. J., Jung, E. J., Park, H., Roy, N., Kim, H., Lee, M. M., Rubin, E. M., Lee, S.-W., & Kim, J. F. (2018). Rhizosphere microbiome structure alters to enable wilt resistance in tomato. Nature Biotechnology, 36(11), 1100–1109. https://doi.org/10.1038/nbt.4232

25

Laforest‐Lapointe, I., & Whitaker, B. K. (2019). Decrypting the phyllosphere microbiota: Progress and challenges. American Journal of Botany, 106(2), 171– 173. https://doi.org/10.1002/ajb2.1229

Lakshmanan, V., Selvaraj, G., & Bais, H. P. (2014). Functional Soil Microbiome: Belowground Solutions to an Aboveground Problem. PLANT PHYSIOLOGY, 166(2), 689–700. https://doi.org/10.1104/pp.114.245811

Lemaire, G., Wilkins, R., & Hodgson, J. (2005). Challenges for grassland science: Managing research priorities. Agriculture, Ecosystems & Environment, 108(2), 99–108. https://doi.org/10.1016/j.agee.2005.01.003

Lennon, J. T., & Jones, S. E. (2011). Microbial seed banks: The ecological and evolutionary implications of dormancy. Nature Reviews Microbiology, 9(2), 119–130. https://doi.org/10.1038/nrmicro2504

Levy, A., Salas Gonzalez, I., Mittelviefhaus, M., Clingenpeel, S., Herrera Paredes, S., Miao, J., Wang, K., Devescovi, G., Stillman, K., Monteiro, F., Rangel Alvarez, B., Lundberg, D. S., Lu, T.-Y., Lebeis, S., Jin, Z., McDonald, M., Klein, A. P., Feltcher, M. E., Rio, T. G., … Dangl, J. L. (2018). Genomic features of bacterial adaptation to plants. Nature Genetics, 50(1), 138–150. https://doi.org/10.1038/s41588-017-0012-9

Liu, B., Tu, C., Hu, S., Gumpertz, M., & Ristaino, J. B. (2007). Effect of organic, sustainable, and conventional management strategies in grower fields on soil physical, chemical, and biological factors and the incidence of Southern blight. Applied Soil Ecology, 37(3), 202–214. https://doi.org/10.1016/j.apsoil.2007.06.007

López, S., Pastorino, G., Franco, M., Medina, R., Lucentini, C., Saparrat, M., & Balatti, P. (2018). Microbial Endophytes that Live within the Seeds of Two Tomato Hybrids Cultivated in Argentina. Agronomy, 8(8), 136. https://doi.org/10.3390/agronomy8080136 Lucas, J. A., García-Cristobal, J., Bonilla, A., Ramos, B., & Gutierrez-Mañero, J. (2014). Beneficial rhizobacteria from rice rhizosphere confers high protection against biotic and abiotic stress inducing systemic resistance in rice seedlings. Plant Physiology and Biochemistry, 82, 44–53. https://doi.org/10.1016/j.plaphy.2014.05.007

Lugtenberg, B., & Kamilova, F. (2009). Plant-growth-promoting rhizobacteria. Annual Review of Microbiology, 63, 541–556. https://doi.org/10.1146/annurev.micro.62.081307.162918

26

Lundberg, D. S., Lebeis, S. L., Paredes, S. H., Yourstone, S., Gehring, J., Malfatti, S., Tremblay, J., Engelbrektson, A., Kunin, V., Del Rio, T. G., Edgar, R. C., Eickhorst, T., Ley, R. E., Hugenholtz, P., Tringe, S. G., & Dangl, J. L. (2012). Defining the core Arabidopsis thaliana root microbiome. Nature, 488(7409), 86– 90. https://doi.org/10.1038/nature11237

Madmony, A., Chernin, L., Pleban, S., Peleg, E., & Riov, J. (2005). Enterobacter cloacae, an obligatory endophyte of pollen grains of Mediterranean pines. Folia Microbiologica, 50(3), 209–216. https://doi.org/10.1007/BF02931568

McFrederick, Q. S., Thomas, J. M., Neff, J. L., Vuong, H. Q., Russell, K. A., Hale, A. R., & Mueller, U. G. (2017). Flowers and Wild Megachilid Bees Share Microbes. Microbial Ecology, 73(1), 188–200. https://doi.org/10.1007/s00248- 016-0838-1

Melotto, M., Underwood, W., & He, S. Y. (2008). Role of Stomata in Plant Innate Immunity and Foliar Bacterial Diseases. Annual Review of Phytopathology, 46(1), 101–122. https://doi.org/10.1146/annurev.phyto.121107.104959

Melotto, M., Underwood, W., Koczan, J., Nomura, K., & He, S. Y. (2006). Plant stomata function in innate immunity against bacterial invasion. Cell, 126(5), 969–980. https://doi.org/10.1016/j.cell.2006.06.054

Mendes, R., Kruijt, M., de Bruijn, I., Dekkers, E., van der Voort, M., Schneider, J. H. M., Piceno, Y. M., DeSantis, T. Z., Andersen, G. L., Bakker, P. A. H. M., & Raaijmakers, J. M. (2011). Deciphering the Rhizosphere Microbiome for Disease-Suppressive Bacteria. Science, 332(6033), 1097–1100. https://doi.org/10.1126/science.1203980

Miliute, I., Buzaite, O., Baniulis, D., & Stanys, V. (2015). Bacterial endophytes in agricultural crops and their role in stress tolerance: A review. Zemdirbyste- Agriculture, 102(4), 465–478. https://doi.org/10.13080/z-a.2015.102.060

Miller, I. M. (1990). Bacterial leaf nodule symbiosis. In Advances in botanical research (Vol. 17, pp. 163–234). Elsevier.

Mueller, U. G., & Sachs, J. L. (2015). Engineering Microbiomes to Improve Plant and Animal Health. Trends in Microbiology, 23(10), 606–617. https://doi.org/10.1016/j.tim.2015.07.009

Müller, D. B., Vogel, C., Bai, Y., & Vorholt, J. A. (2016). The Plant Microbiota: Systems-Level Insights and Perspectives. Annual Review of Genetics, 50(1), 211–234. https://doi.org/10.1146/annurev-genet-120215-034952

27

Mundt, J. O., & Hinkle, N. F. (1976). Bacteria within ovules and seeds. Applied and Environmental Microbiology, 32(5), 694–698.

Naveed, M., Mitter, B., Yousaf, S., Pastar, M., Afzal, M., & Sessitsch, A. (2014). The endophyte Enterobacter sp. FD17: A maize growth enhancer selected based on rigorous testing of plant beneficial traits and colonization characteristics. Biology and Fertility of Soils, 50(2), 249–262. https://doi.org/10.1007/s00374- 013-0854-y

Nelson, E. B. (2004). MICROBIAL DYNAMICS AND INTERACTIONS IN THE SPERMOSPHERE. Annual Review of Phytopathology, 42(1), 271–309. https://doi.org/10.1146/annurev.phyto.42.121603.131041

Newcombe, G., Harding, A., Ridout, M., & Busby, P. E. (2018). A Hypothetical Bottleneck in the Plant Microbiome. Frontiers in Microbiology, 9. https://doi.org/10.3389/fmicb.2018.01645

Oldroyd, G. E. D. (2013). Speak, friend, and enter: Signalling systems that promote beneficial symbiotic associations in plants. Nature Reviews. Microbiology, 11(4), 252–263. https://doi.org/10.1038/nrmicro2990

Ottesen, A. R., González Peña, A., White, J. R., Pettengill, J. B., Li, C., Allard, S., Rideout, S., Allard, M., Hill, T., Evans, P., Strain, E., Musser, S., Knight, R., & Brown, E. (2013). Baseline survey of the anatomical microbial ecology of an important food plant: Solanum lycopersicum (tomato). BMC Microbiology, 13(1), 114. https://doi.org/10.1186/1471-2180-13-114

Parniske, M. (2008). Arbuscular mycorrhiza: The mother of plant root endosymbioses. Nature Reviews Microbiology, 6(10), 763–775. https://doi.org/10.1038/nrmicro1987

Pavlo, A., Leonid, O., Iryna, Z., Natalia, K., & Maria, P. A. (2011). Endophytic bacteria enhancing growth and disease resistance of potato (Solanum tuberosum L.). Biological Control, 56(1), 43–49. https://doi.org/10.1016/j.biocontrol.2010.09.014

Peiffer, J. A., Spor, A., Koren, O., Jin, Z., Tringe, S. G., Dangl, J. L., Buckler, E. S., & Ley, R. E. (2013). Diversity and heritability of the maize rhizosphere microbiome under field conditions. Proceedings of the National Academy of Sciences, 110(16), 6548–6553. https://doi.org/10.1073/pnas.1302837110

Pellegrino, E., Turrini, A., Gamper, H. A., Cafà, G., Bonari, E., Young, J. P. W., & Giovannetti, M. (2012). Establishment, persistence and effectiveness of arbuscular mycorrhizal fungal inoculants in the field revealed using molecular 28

genetic tracing and measurement of yield components. New Phytologist, 194(3), 810–822. https://doi.org/10.1111/j.1469-8137.2012.04090.x

Peralta, A. L., Sun, Y., McDaniel, M. D., & Lennon, J. T. (2018). Crop rotational diversity increases disease suppressive capacity of soil microbiomes. Ecosphere, 9(5), e02235. https://doi.org/10.1002/ecs2.2235

Podile, A. R., Vukanti, R., Ankati, S., Kalam, S., Dutta, S., Durgeshwar, P., & Rao, V. (2014). Root Colonization and Quorum Sensing are the Driving forces of Plant Growth Promoting Rhizobacteria (pgpr) for Growth Promotion. Proceedings of the Indian National Science Academy, 80, 407. https://doi.org/10.16943/ptinsa/2014/v80i2/55117

Pradhanang, P. M., Elphinstone, J. G., & Fox, R. T. V. (2000). Identification of crop and weed hosts of Ralstonia solanacearum biovar 2 in the hills of Nepal. Plant Pathology, 49(4), 403–413. https://doi.org/10.1046/j.1365-3059.2000.00480.x

Raynaud, X., & Nunan, N. (2014). Spatial Ecology of Bacteria at the Microscale in Soil. PLoS ONE, 9(1). https://doi.org/10.1371/journal.pone.0087217

Reinhold-Hurek, B., Maes, T., Gemmer, S., Van Montagu, M., & Hurek, T. (2006). An Endoglucanase Is Involved in Infection of Rice Roots by the Not-Cellulose- Metabolizing Endophyte Azoarcus Sp. Strain BH72. Molecular Plant-Microbe Interactions®, 19(2), 181–188. https://doi.org/10.1094/MPMI-19-0181

Remus‐Emsermann, M. N. P., Lücker, S., Müller, D. B., Potthoff, E., Daims, H., & Vorholt, J. A. (2014). Spatial distribution analyses of natural phyllosphere- colonizing bacteria on Arabidopsis thaliana revealed by fluorescence in situ hybridization. Environmental Microbiology, 16(7), 2329–2340. https://doi.org/10.1111/1462-2920.12482

Rho, H., Van Epps, V., Wegley, N., Doty, S. L., & Kim, S.-H. (2018). Salicaceae Endophytes Modulate Stomatal Behavior and Increase Water Use Efficiency in Rice. Frontiers in Plant Science, 9. https://doi.org/10.3389/fpls.2018.00188 Roberts, D. P., Baker, C. J., McKenna, L., Liu, S., Buyer, J. S., & Kobayashi, D. Y. (2009). Influence of host seed on metabolic activity of Enterobacter cloacae in the spermosphere. Soil Biology and Biochemistry, 41(4), 754–761. https://doi.org/10.1016/j.soilbio.2009.01.010

Rosenblueth, M., & Martínez-Romero, E. (2006). Bacterial Endophytes and Their Interactions with Hosts. Molecular Plant-Microbe Interactions®, 19(8), 827– 837. https://doi.org/10.1094/MPMI-19-0827

29

Rudrappa, T., Czymmek, K. J., Paré, P. W., & Bais, H. P. (2008). Root-Secreted Malic Acid Recruits Beneficial Soil Bacteria. Plant Physiology, 148(3), 1547–1556. https://doi.org/10.1104/pp.108.127613

Saile, E., McGarvey, J. A., Schell, M. A., & Denny, T. P. (1997). Role of Extracellular Polysaccharide and Endoglucanase in Root Invasion and Colonization of Tomato Plants by Ralstonia solanacearum. Phytopathology®, 87(12), 1264– 1271. https://doi.org/10.1094/PHYTO.1997.87.12.1264

Santhanam, R., Luu, V. T., Weinhold, A., Goldberg, J., Oh, Y., & Baldwin, I. T. (2015). Native root-associated bacteria rescue a plant from a sudden-wilt disease that emerged during continuous cropping. Proceedings of the National Academy of Sciences, 112(36), E5013–E5020. https://doi.org/10.1073/pnas.1505765112

Sapers, G. M., Gorny, J. R., & Yousef, A. E. (2005). Microbiology of Fruits and Vegetables. CRC Press.

Scharf, B. E., Hynes, M. F., & Alexandre, G. M. (2016). Chemotaxis signaling systems in model beneficial plant–bacteria associations. Plant Molecular Biology, 90(6), 549–559. https://doi.org/10.1007/s11103-016-0432-4

Schiltz, S., Gaillard, I., Pawlicki‐Jullian, N., Thiombiano, B., Mesnard, F., & Gontier, E. (2015). A review: What is the spermosphere and how can it be studied? Journal of Applied Microbiology, 119(6), 1467–1481. https://doi.org/10.1111/jam.12946 Schlatter, D. C., Bakker, M. G., Bradeen, J. M., & Kinkel, L. L. (2015). Plant community richness and microbial interactions structure bacterial communities in soil. Ecology, 96(1), 134–142. https://doi.org/10.1890/13-1648.1

Schmidt, J. E., Kent, A. D., Brisson, V. L., & Gaudin, A. C. M. (2019). Agricultural management and plant selection interactively affect rhizosphere microbial community structure and nitrogen cycling. Microbiome, 7(1), 146. https://doi.org/10.1186/s40168-019-0756-9

Schulz-Bohm, K., Gerards, S., Hundscheid, M., Melenhorst, J., Boer, W. de, & Garbeva, P. (2018). Calling from distance: Attraction of soil bacteria by plant root volatiles. The ISME Journal, 12(5), 1252–1262. https://doi.org/10.1038/s41396-017-0035-3

Scott, R. I., Chard, J. M., Hocart, M. J., Lennard, J. H., & Graham, D. C. (1996). Penetration of potato tuber lenticels by bacteria in relation to biological control of blackleg disease. Potato Research, 39(3), 333–344. https://doi.org/10.1007/BF02357937

30

Shahzad, R., Khan, A. L., Bilal, S., Asaf, S., & Lee, I.-J. (2018). What Is There in Seeds? Vertically Transmitted Endophytic Resources for Sustainable Improvement in Plant Growth. Frontiers in Plant Science, 9. https://doi.org/10.3389/fpls.2018.00024

Shaik, S. P., & Thomas, P. (2019). In Vitro Activation of Seed-Transmitted Cultivation- Recalcitrant Endophytic Bacteria in Tomato and Host–Endophyte Mutualism. Microorganisms, 7(5). https://doi.org/10.3390/microorganisms7050132

Simon, J.-C., Marchesi, J. R., Mougel, C., & Selosse, M.-A. (2019). Host-microbiota interactions: From holobiont theory to analysis. Microbiome, 7(1), 5. https://doi.org/10.1186/s40168-019-0619-4

Sinnesael, A., Eeckhout, S., Janssens, S. B., Smets, E., Panis, B., Leroux, O., & Verstraete, B. (2018). Detection of Burkholderia in the seeds of Psychotria punctata (Rubiaceae) – Microscopic evidence for vertical transmission in the leaf nodule symbiosis. PLOS ONE, 13(12), e0209091. https://doi.org/10.1371/journal.pone.0209091

Smalla, K., Wieland, G., Buchner, A., Zock, A., Parzy, J., Kaiser, S., Roskot, N., Heuer, H., & Berg, G. (2001). Bulk and Rhizosphere Soil Bacterial Communities Studied by Denaturing Gradient Gel Electrophoresis: Plant-Dependent Enrichment and Seasonal Shifts Revealed. Applied and Environmental Microbiology, 67(10), 4742–4751. https://doi.org/10.1128/AEM.67.10.4742- 4751.2001

Souza, R. de, Ambrosini, A., Passaglia, L. M. P., Souza, R. de, Ambrosini, A., & Passaglia, L. M. P. (2015). Plant growth-promoting bacteria as inoculants in agricultural soils. Genetics and Molecular Biology, 38(4), 401–419. https://doi.org/10.1590/S1415-475738420150053

Srour, A. Y., Gibson, D. J., Leandro, L. F. S., Malvick, D. K., Bond, J. P., & Fakhoury, A. M. (2017). Unraveling Microbial and Edaphic Factors Affecting the Development of Sudden Death Syndrome in Soybean. Phytobiomes Journal, 1(2), 91–101. https://doi.org/10.1094/PBIOMES-02-17-0009-R

Sun, H. Y., Deng, S. P., & Raun, W. R. (2004). Bacterial Community Structure and Diversity in a Century-Old Manure-Treated Agroecosystem. Applied and Environmental Microbiology, 70(10), 5868–5874. https://doi.org/10.1128/AEM.70.10.5868-5874.2004

Sun, Y., Cheng, Z., & Glick, B. R. (2009). The presence of a 1-aminocyclopropane-1- carboxylate (ACC) deaminase deletion mutation alters the physiology of the

31

endophytic plant growth-promoting bacterium Burkholderia phytofirmans PsJN. FEMS Microbiology Letters, 296(1), 131–136.

Tallman, G. (2004). Are diurnal patterns of stomatal movement the result of alternating metabolism of endogenous guard cell ABA and accumulation of ABA delivered to the apoplast around guard cells by transpiration? Journal of Experimental Botany, 55(405), 1963–1976. https://doi.org/10.1093/jxb/erh212

Thomas, P., Kumari, S., Swarna, G. K., & Gowda, T. K. S. (2007). Papaya shoot tip associated endophytic bacteria isolated from in vitro cultures and host- endophyte interaction in vitro and in vivo. Canadian Journal of Microbiology, 53(3), 380–390. https://doi.org/10.1139/W06-141

Thomas, P., & Sekhar, A. C. (2014). Live cell imaging reveals extensive intracellular cytoplasmic colonization of banana by normally non-cultivable endophytic bacteria. AoB PLANTS, 6(0). https://doi.org/10.1093/aobpla/plu002

Togashi, J., Ueda, K., & Namai, T. (2001). Overwintering of Erwinia carotovora subsp. Carotovora in Diseased Tissues in Soil and Its Role as Inoculum for Soft Rot of Chinese Cabbage (Brassica campestris, Pekinensis Group). Journal of General Plant Pathology, 67(1), 45–50. https://doi.org/10.1007/PL00012986

Truyens, S., Weyens, N., Cuypers, A., & Vangronsveld, J. (2015). Bacterial seed endophytes: Genera, vertical transmission and interaction with plants. Environmental Microbiology Reports, 7(1), 40–50. https://doi.org/10.1111/1758-2229.12181

Turner, T. R., James, E. K., & Poole, P. S. (2013). The plant microbiome. Genome Biology, 14(6), 209. https://doi.org/10.1186/gb-2013-14-6-209

Udvardi, M., & Poole, P. S. (2013). Transport and Metabolism in Legume-Rhizobia Symbioses. Annual Review of Plant Biology, 64(1), 781–805. https://doi.org/10.1146/annurev-arplant-050312-120235

Ushio, M., Yamasaki, E., Takasu, H., Nagano, A. J., Fujinaga, S., Honjo, M. N., Ikemoto, M., Sakai, S., & Kudoh, H. (2015). Microbial communities on flower surfaces act as signatures of pollinator visitation. Scientific Reports, 5(1), 1–7. https://doi.org/10.1038/srep08695

Vendan, R. T., Yu, Y. J., Lee, S. H., & Rhee, Y. H. (2010). Diversity of endophytic bacteria in ginseng and their potential for plant growth promotion. The Journal of Microbiology, 48(5), 559–565. https://doi.org/10.1007/s12275-010-0082-1

32

Voort, M. van der, Kempenaar, M., Driel, M. van, Raaijmakers, J. M., & Mendes, R. (2016). Impact of soil heat on reassembly of bacterial communities in the rhizosphere microbiome and plant disease suppression. Ecology Letters, 19(4), 375–382. https://doi.org/10.1111/ele.12567

Wakelin, S. A., Barratt, B. I. P., Gerard, E., Gregg, A. L., Brodie, E. L., Andersen, G. L., DeSantis, T. Z., Zhou, J., He, Z., Kowalchuk, G. A., & O’Callaghan, M. (2013). Shifts in the phylogenetic structure and functional capacity of soil microbial communities follow alteration of native tussock grassland ecosystems. Soil Biology and Biochemistry, 57, 675–682. https://doi.org/10.1016/j.soilbio.2012.07.003

Walker, T. S., Bais, H. P., Grotewold, E., & Vivanco, J. M. (2003). Root Exudation and Rhizosphere Biology. Plant Physiology, 132(1), 44–51. https://doi.org/10.1104/pp.102.019661

Wang, Z., Liu, L., Chen, Q., Wen, X., & Liao, Y. (2016). Conservation tillage increases soil bacterial diversity in the dryland of northern China. Agronomy for Sustainable Development, 36(2), 28. https://doi.org/10.1007/s13593-016-0366-x

Wardhaugh, C. W., Stork, N. E., Edwards, W., & Grimbacher, P. S. (2012). The Overlooked Biodiversity of Flower-Visiting Invertebrates. PLOS ONE, 7(9), e45796. https://doi.org/10.1371/journal.pone.0045796

Wardle, D. A., Bardgett, R. D., Klironomos, J. N., Setälä, H., Putten, W. H. van der, & Wall, D. H. (2004). Ecological Linkages Between Aboveground and Belowground Biota. Science, 304(5677), 1629–1633. https://doi.org/10.1126/science.1094875

Wei, G. (1996). Induced Systemic Resistance to Cucumber Diseases and Increased Plant Growth by Plant Growth-Promoting Rhizobacteria Under Field Conditions. Phytopathology, 86(2), 221. https://doi.org/10.1094/Phyto-86-221

Weller, D. M., Raaijmakers, J. M., Gardener, B. B. M., & Thomashow, L. S. (2002). Microbial Populations Responsible For Specific Soil Suppressiveness To Plant Pathogens. Annual Review of Phytopathology, 40(1), 309–348. https://doi.org/10.1146/annurev.phyto.40.030402.110010

Zhao, Y., Gao, Z., Tian, B., Bi, K., Chen, T., Liu, H., Xie, J., Cheng, J., Fu, Y., & Jiang, D. (2017). Endosphere microbiome comparison between symptomatic and asymptomatic roots of Brassica napus infected with Plasmodiophora brassicae. PLOS ONE, 12(10), e0185907. https://doi.org/10.1371/journal.pone.0185907

33

Zhou, J. Y., Zhao, X. Y., & Dai, C. C. (2014). Antagonistic mechanisms of endophytic Pseudomonas fluorescens against Athelia rolfsii. Journal of Applied Microbiology, 117(4), 1144–1158. https://doi.org/10.1111/jam.12586

Zinniel, D. K., Lambrecht, P., Harris, N. B., Feng, Z., Kuczmarski, D., Higley, P., Ishimaru, C. A., Arunakumari, A., Barletta, R. G., & Vidaver, A. K. (2002). Isolation and Characterization of Endophytic Colonizing Bacteria from Agronomic Crops and Prairie Plants. Applied and Environmental Microbiology, 68(5), 2198–2208. https://doi.org/10.1128/AEM.68.5.2198-2208.2002

34

Chapter 2. Characterization of the Bacterial Endophytic Microbiome of Tomato

Seedlings Grown in a Soil Diversity Gradient Using Chicken-Grazed Soils

Introduction

Endophytic bacteria reside inside plant organs and are considered important components of the plant microbiome (Bacon & White, 2000). The establishment of endophytic and rhizosphere microbiome is influenced mainly by the plant host and soil.

Many bacterial endophytes are attracted by root exudates and rhizodeposits, enabling their colonization of plant hosts through the roots (Kawasaki et al., 2016). Other areas of plant entry include stem and leaf, surfaces, flowers, fruits, stomata, wounds and hydathodes (Compant et al., 2010). These bacteria have exhibited functional, as well as taxonomic diversity. Bacterial endophytes with plant growth promoting activities can positively impact host growth through indole acetic acid production, nitrogen fixation and phosphate solubilization (Ma et al., 2011, 2016; Sexton et al., 2020; Tariq et al., 2014).

Additionally, these bacteria can also engage in antagonistic activity with plant pathogens and promote disease suppression (Shahzad et al., 2017; Xu et al., 2019; Yanti et al.,

2019). Proteobacteria, Actinobacteria, Firmicutes, and Bacteroidetes are dominant across plant environments, harboring bacteria with diverse ecological relationships ranging from mutualists to pathogenic (Cordero et al., 2020; Deng et al.,

2019; Szymańska et al., 2018; Wang et al., 2019). Some of these bacteria can also be 35 pathogens to humans while providing benefits to the plant host (Chelius & Triplett,

2000; E. T. Wang et al., 2006).

Diverse forces drive the multitrophic interactions surrounding said bacterial endophyte establishment and composition. Soil type and host plant genotype has been observed to shape bacterial endophyte community composition, with soil type possibly being a stronger driver (Ding et al., 2013; Pablo R. Hardoim et al., 2015; Rasche et al.,

2006). Furthermore, plant growth stage and agricultural management strategies can shift community composition (Gdanetz & Trail, 2017; Ren et al., 2015; Xia et al., 2015).

Crop-livestock systems are gaining interest due to greater energy efficiency, nutrient recycling, perenniality and opportunity to diversify rotations (McGilloway, 2005).

Integrated crop-livestock systems are a type of mixed farming practice where farmers can have different enterprises and save resources by interchanging them on the farm (Schiere

& Kater, 2001). In crop-livestock systems, small farmers can cash income, improve quality and quantity of food produced and exploitation of unutilized resources by using the byproduct of one system as the input for other (Gupta et al., 2012). It has been found that crop-livestock systems can increase soil quality and fertility, reduce cost of herbicide use and improve sustainability, especially for farmers in poorer areas of the world (de

Faccio Carvalho et al. 2010; Devendra and Thomas, 2002). Studies involving the impact of this type of management strategies on microbial communities are not common. Among these few studies, one showed an increase in microbial biomass and enzyme activities in soil samples from a crop-livestock system compared to a monoculture (Acosta-Martínez

36 et al. 2010). Others have shown the enhancement of soil fertility, productivity and bioremediation with chicken manure applications (Dikinya and Mufwanzala, 2010; Liu et al. 2009). Additionally, crop-livestock systems include the use of cover crops or forage which provide a grazing source for animals (Gardner et al., 1991; Luo et al., 2017). As cover crops increase plant diversity, they also have positive effects on microbial biomass

(Chen et al., 2019; Michalk et al., 2017). Although crop-livestock systems have demonstrated their efficiency and improved sustainability, the impact of this strategy on the endophytic plant microbiome is not completely known.

To further research microbiomes and account for the various forces driving bacterial community composition, it is necessary to establish simplified model plant- microbiome systems. Conducting controlled experiments can assist in streamlining plant- microbiome interactions and aid in uncovering taxonomic and functional diversity of microbes (Chihaoui et al., 2015; Pavlova et al., 2017; Wicaksono et al., 2018). Various factors in these experiments are modified, such as growth conditions, planting substrate type and quality, temperature, and duration of experiment. Tomato plants have already been used to obtain insight on the role of endophytes and the microbial composition of different plant tissues (Romero et al., 2014; Tian et al., 2015). Other than being one of the top ten most important crops in the world, the tomato plant’s short lifecycle and ability to adapt to different conditions, makes it an appropriate research model for our purposes

(Bergougnoux, 2014). By simplifying interactions, greater understanding can be obtained from the impact of agricultural management strategies on crops and associated microbes.

37

Advances in next-generation sequencing, alongside traditional culturing approaches, have facilitated recovery, detection and identification of endophytes

(Quambusch & Winkelmann, 2018). Although it is often time consuming, culturing bacteria remains an important step to understand their biological processes and ecological roles as well as obtaining novel insights for application (Palleroni, 1997). Coupled with advances in culture-independent approaches, information provided by obtaining bacterial isolates presents the opportunity to delve into unknown biosynthetic pathways and biochemical characteristics as well as understanding adaptations towards environmental changes (Connon & Giovannoni, 2002; Overmann et al., 2017). However, not meeting all the nutritional requirements of bacteria present creates a bias. For endophytic bacteria, two of the most commonly used media are tryptic soy agar (TSA) and Reasoner's 2A

Agar (R2A). The former supports growth of a broad range of bacteria and the latter targets bacteria that require a lower amount of nutrients (J. M. Gardner et al., 1982;

Reasoner & Geldreich, 1985). It is not suggested to use TSA at full strength due to its high nutrient concentration which allows fast growing bacteria to overgrow those that are slower growing (Hallmann et al., 2006). The bacterial ribosomal marker 16S rRNA gene region is used to identify bacteria at a genus level. Because of its functional stability, high conservation, and its ubiquitous nature, the 16S rRNA gene is an effective molecular marker for taxonomic purposes (Ramasamy et al., 2014). The rpoB gene region has been suggested as an alternative, as well as a compliment, to the 16S rRNA gene due to its comparable phylogenetic resolution (Case et al., 2007). Additionally, the available databases, such as National Center of Biotechnology Information (NCBI) nucleotide

38 database and the Ribosomal Database Project (RDP), assist in sequence identification

(Altschul et al., 1990; Wang et al., 2007).

The aim of this chapter is to address the impact of diluted chicken-grazed soil, as well as the associated bacterial microbiome, on tomato seedling growth and stem endophyte composition. Diluting soil from the field in a greenhouse setting will assist in reducing the impact of a subset of driving forces in bacterial endophyte community composition and plant growth. Driving forces, such as soil’s available nutrients, physical and chemical characteristics, and abundance of microorganisms, can be affected by dilution. The objectives are as follows: (a) determine the impact of chicken-grazed soil dilution treatments on aboveground biomass and height of tomato seedlings, (b) determine the impact of diluted chicken-grazed soil slurry on aboveground and root biomass, number of compound leaves and height of each tomato seedling and (c) evaluate the impact of dilution treatments on cultured endophytic bacteria composition obtained from the stem of tomato seedlings. We hypothesize that as planting substrate contains higher amounts of chicken-grazed soil, tomato seedlings will have higher plant biomass, as well as bacterial endophyte diversity. Furthermore, bacterial endophytes recovered from tomato seedlings grown in planting substrate containing chicken-grazed soil will be similar to those commonly found in plants, and possibly to those associated to animal sources, such as chicken manure. This project will help understand the potential of integrating crop-livestock strategies from the perspective of tomato endophyte diversity and beneficial effects in tomato plant growth.

39

Materials and Methods

Two experiments were conducted in the greenhouse with the objective to determine the impact of diluted chicken-grazed soils on tomato seedling growth and stem endophyte composition. Even though the goals were similar, different approaches were used in 2018 and 2020 and are described below.

Field Site Description

In the year 2016, the project Diversification of Small and Medium Sized Farms

Grant (SMSFG) was established at Mellinger Research Farm at the Ohio Agricultural

Research and Development Center (OARDC), The Ohio State University (OSU). The

0.61-hectare experimental field includes integrated crop and livestock systems as part of a rotation and its impact is evaluated on subsequent vegetable yield (Figure 2.1). Each plot undergoes a management strategy with rotations every year (Table 2.1). For this dissertation, we focused on the pasture and chicken grazing rotation (split plot pasture and pasture chicken). The pasture plots were planted with a mixture of grass and clover for first year establishment and were mowed prior to the growing season for the second- year pasture. The composition of the seed mixture for pasture plots was 13.61 kg of HQ-

R Pasture Mix (JDS Seeds Orrville, OH) with 2.27 kg of white (JDS Seeds Orrville, OH) and red clover (RKO). HQ-R Pasture Mix contains 29% Niva orchard grass, 29% Laura meadow fescue, 20% premium perennial rye grass, 10% Calibra perennial rye grass and

40

10% Climax timothy. This seed mixture was added to new pasture plots established in

2017, 2018 and 2019. In August 2018 and 2019, broiler chickens were placed in tractors and moved across selected pasture split plots for 32 and 35 days, respectively. Chickens grazed the mixture of available pasture and chicken diet was supplemented with feed.

Soil Sampling

2018 Experiment. In October, 20 soil cores were collected from four chicken grazed pasture plots (CGPPs) R1r5, R4r3, R2r1 and R3r1 (Figure 2.1, Table 2.1), approximately ~ four weeks after chickens were removed from the plots, using a 1” soil core sampler and cleaning core sampler with ethanol between plots. Soil cores were bagged in separate plastic bags per plot. All soil recovered from CGPPs were mixed- together and homogenized.

2020 Experiment. In January 29 approximately ~ 20 weeks after chickens were removed from the plots), five soil cores were collected from four CGPPs (R4r2, R3r2,

R4r4, R3r5 (Figure 2.1, Table 2.1)) using a 1” soil core sampler and mixed in one plastic bag. A week later, five soil cores were collected from the same CGPPs using a 1” soil core sampler and mixed in one plastic bag. As in 2018, all soil recovered from CGPPs were mixed-together and homogenized.

41

Planting Substrate, Plant Materials and Growth Conditions

2018 Experiment. Five treatments were generated by diluting homogenized soil from CGPPs with autoclaved soil substrate: 100% chicken-grazed soil, 75%, 50% and

25% dilutions and 100% autoclaved soil substrate (Figure 2.2).

The soil substrate was prepared by mixing 0.0816 metric tons of field soil, 0.0272 metric tons of peat moss, 454 g of lime and top fertilizer, and was autoclaved two times at 121oC. Tomato seeds were sown in 10.16 cm plastic pots filled with soil from each dilution treatment (Figure 2.2). One tomato seed was sown for each pot. Tomato seeds from two varieties were used: Celebrity (Seeds from September 11, 2018 originally from tomatoes grown in CGPPs at Mellinger Research Farm Plot R3r3) and Moneymaker

(Everwilde Farms), with five pots per tomato variety per treatment for a total of 50 pots.

The experiment was setup as a randomized complete block design, where each block represented a different tomato variety. Pots were evaluated for germination 12 days after sowing and no significant differences in seed germination were observed. Greenhouse conditions ranged from a maximum of 25oC and a minimum of 21oC.

2020 Experiment. To develop a controlled inoculation, while reducing the impact of soil’s available nutrients, physical and chemical characteristics, and abundance of microorganisms, a chicken-grazed soil slurry was prepared. For this, 60g of soil obtained from CGPPs was mixed with 400mL of sterile water in a sterile 1L bottle for 10 minutes,

42 creating a soil slurry. The soil slurry was passed through a stainless-steel filter with 47 mm pore size, disinfected with bleach, using a vacuum pump. 10 mL of the collected filtered soil slurry was added to a sterile beaker with 100mL of water, diluting it to 10-1.

The dilution was repeated three times generating four soil slurry treatments: 10-1, 10-2, 10-

3 and 10-4 (Figure 2.3). Sterile water was used as a control.

To measure the concentration of colony forming units (CFUs) of each slurry at the time of application, 100 µL of each dilution, including the control, was spread on a

1/10 TSA plate and colonies were counted after 24 hours. Seventy-five Celebrity F1 tomato seeds (Johnny’s Seeds) were surface sterilized in 0.5% NaOCl and 0.1% Tween for 1 minute and washed using sterile water. For surface sterilization, tomato seeds were contained in a plastic tissue cassette (Tissue Path, Thermo Scientific). For each treatment, autoclaved germination paper was moistened inside the beaker. Approximately 20 surface-sterilized seeds were placed 2.54 cm apart on one piece of autoclaved germination paper for each treatment, including control. To facilitate the establishment of microorganisms originating from chicken-grazed soils, tomato seeds were inoculated with diluted soil slurry treatments, prior to germination. Papers were rolled and placed in separate beakers for each treatment. Each beaker was covered with aluminum foil and placed in a growing chamber with lights on for 16 hours of the day for one week. Six days later, seedlings were transferred to pots containing a growing substrate of sand, turface, and composted soil. The substrate was prepared by mixing sand:turface 1:1 followed by a final mixture of 2/3 sand-turface and 1/3 composted soil. The planting

43 substrate was autoclaved two times at 121oC. Fifty 10.16 cm pots were sprayed with ethanol and filled with planting substrate. Pots were watered and after 24 hours two pre- germinated tomato seedlings were sowed per pot. By using newly sampled chicken- grazed soil, a second soil slurry treatment was prepared, filtered and diluted using the same method as previously described. A total of five treatments were prepared: 10-1, 10-2,

10-3 and 10-4, including sterile water as a control. After sowing seedlings, 30ml of each treatment was drenched on pots corresponding to the original treatment of the pre- germinated seeds. Pots were placed in a randomized block design and grown for six weeks in the greenhouse. After a week, seedlings in each pot were thinned to one. After two weeks, 10ml of NPK (20-20-20) fertilizer was added to each pot once a week.

Greenhouse conditions ranged from a maximum average of 25oC and a minimum average of 23oC.

Seedling Sampling and Evaluation

2018 Experiment. Seedlings were sampled at the 4-leaf stage, approximately six weeks after sowing. Seedlings had the aboveground plant tissues separated from the below-ground by cutting with scissors. Roots were rinsed to remove planting substrate.

Aboveground plant tissues were measured for biomass and height.

2020 Experiment. Seedlings were sampled at the 4-leaf stage, approximately six weeks after sowing. Seedling had the aboveground plant tissues separated from the

44 below-ground by cutting with scissors. Roots were rinsed to remove planting substrate.

Aboveground plant tissues were measured for biomass, height and number of compound leaves. Root biomass was also weighed.

Plant Tissue Sampling and Isolation of Endophytic Bacteria from Tomato Seedling Stem

For both 2018 and 2020 experiments, stems were surface sterilized through a stepwise procedure by immersing tissue in 70% ethanol (1 min), 20% bleach mixed with

0.1% Tween (1 min), 70% ethanol (1 min) and rinsed three times with sterile distilled water. Edges of the surface sterilized stem were cut off. Surface-sterilized stem was cut into several 1-cm section pieces. Sections were plated on 1/10 TSA and R2A and incubated at ambient temperature (approximately 22oC) for 48 hours. Emerging bacterial colonies from the stem sections were transferred up to five times to obtain pure cultures.

DNA Extraction, PCR Amplification and Sequencing of the 16S rRNA and rpoB Gene Regions from Individual Bacterial Isolates

To characterize bacterial endophytes recovered, molecular techniques were employed. A boiling lysis method, according to Sanders and Miller (2010), was used to extract genomic DNA from individual bacterial isolates. A single colony was selected and immersed in 200µL of sterile water in a 1.5mL tube. The tube was then incubated in a water bath at 100oC for 10 minutes, transferred to ice for 10 minutes followed by

45 centrifugation at 10,000 rpm for 10 minutes. 150µL of the supernatant containing the genomic DNA was transferred to a new 1.5mL tube. PCR amplification of the 16SrRNA gene for bacterial endophytes was performed using primers 8F (5’-

TGGAGAGTTTGATCCTGGCTCAG-3’) and 1492R (5'-

GGTTACCTTGTTACGACTT-3') (Weisburg et al., 1991).

The PCR reaction mixture consisted of 2 μL of DNA template, 1× GoTaq Flexi

Buffer (Promega), 0.2 mM dNTP mix (Promega), 2 mM of MgCl2 (Promega), 1 mg/ml of bovine serum albumin (BSA), 0.03 GoTaq DNA polymerase (Promega), 0.5 μM each primer, and molecular grade water (Hyclone). PCR reaction conditions were an initial denaturation at 95oC for 5 minutes, 35 cycles of denaturation at 94oC for 1 minute, annealing at 54oC for 45 seconds and elongation at 72oC for 1 minute and a final elongation cycle of 72oC for 8 minutes. 5µL of each amplification product was analyzed by 1% agarose gel (1%, containing 5 μL of 10,000X SYBR Safe, Thermo Fisher

Scientific) electrophoresis in 0.5 Tris borate EDTA (TBE) buffer with 100 bp Plus DNA ladder (Promega).

To identify recovered isolates at a species level, amplification of the rpoB gene region was performed. The rpoB region was selected to be amplified due to it exhibiting higher levels of divergence compared to 16S rRNA gene region (Mollet et al., 1997; Vos et al., 2012). First, bacterial genomic DNA was purified using Wizard Genomic DNA

Purification Kit (Promega). PCR amplification of the rpoB gene for bacterial endophytes

46 of the genus Bacillus was performed using primers rpoB‐f (5′‐

AGGTCAACTAGTTCAGTATGGAC‐3′) and rpoB‐r (5′‐

AAGAACCGTAACCGGCAACTT‐3′). The PCR reaction mixture of 25 μL for rpoB genes using rpoB-f and rpoB-r primer pair was adapted from Özdemir & Arslan

(2019), which included 2 μL of DNA template, 1× GoTaq Flexi Buffer (Promega), 0.2 mM dNTP mix (Promega), 2 mM of MgCl2 (Promega), 1 mg/ml of BSA, 0.03 GoTaq

DNA polymerase (Promega), 0.4 μM each primer, and molecular grade water (Hyclone).

PCR amplification of the rpoB gene for the remaining bacterial endophytes was performed using primers Univ_rpoB-F_deg (5′‐GGYTWYGAAGTNCGHGACGTDCA‐

3′) and Univ_rpoB_R_deg (5′‐TGACGYTGCATGTTBGMRCCCATMA‐3′) (Ogier et al., 2019). The PCR reaction mixture of 25 μL for rpoB genes using Univ_rpoB-F_deg and Univ_rpoB_R_deg primer pair included 3 μL of DNA template, 1× GoTaq Flexi

Buffer (Promega), 0.2 mM dNTP mix (Promega), 1.5 mM of MgCl2 (Promega), 1 mg/ml of BSA, 0.03 GoTaq DNA polymerase (Promega), 0.2 μM each primer, and molecular grade water (Hyclone). The PCR reaction conditions for both primer pairs were adapted from Clerck and Vos (2004) and it consisted of denaturation at 94°C for 5 minutes, 40 cycles of denaturation at 94°C for 30 seconds, annealing at 53°C for 45 seconds and elongation at 72°C for 1 minute with a final elongation cycle at 72°C for 5 minutes. Two additional rpoB primer pairs were tested for PCR amplification: rpoB-985-F (5’-

ACNGAYATGATMAAGGG-3’) and rpoB-2362-R (5’-GCCATYARMCCSCGCAT-3’) and RpoB-F (5’-AACCAGTTCCGCGTTGGCCTGG-3’) and RpoB-R (5’-

47

CCTGAACAACACGCTCGGA-3’) (Hoffmann & Roggenkamp, 2003; Jiao, 2013).

These were not successful in amplifying the isolate collection from tomato.

PCR products were evaluated through gel electrophoresis. The gels were visualized under UV transillumination (Gel Logic 112 Imaging System, Kodak).

Enzymatic PCR cleanup was done by adding 5µL of PCR product, 1µL of Alkaline

Phosphatase and 1µL Exonuclease I (Ilustra) into a new 0.2ml tube. PCR cleanup conditions consisted incubating at 37°C for 15 minutes followed by incubating at 80°C for 15 minutes. Preparation of cleaned PCR products for sequencing consisted of adding

10.75µL of clean PCR product and 1.25µL of 6.4 pmol of one of two primers previously used for PCR reaction. Samples were sequenced at the Genomics Shared Resource at The

Ohio State University Comprehensive Cancer Center, Columbus, OH. Partial nucleotide sequence was determined with Sanger sequencing using forward and reverse primers.

Isolate Sequence Data Analysis

To evaluate the sequences, chromatograms were visually revised and trimmed for quality control using 4Peaks software (Gerberastraat & Aalsmeer, 2001). Sequenced ends were trimmed based on quality. Forward and reverse reads for each isolate were merged with CLC Sequence Viewer (Genomics Workbench Version 20.0, 2020). Sequences were compared to the NCBI (National Center of Biotechnology Information) nucleotide database using the basic local alignment search tool (BLAST, blastn) (Altschul et al.,

48

1990), and to the Ribosomal Database Project (RDP) (Wang et al., 2007) using Classifier to classify sequences to genus. Phylogenetic analysis was performed using MEGA software version 7 (Kumar et al., 2016). Recovered isolate sequences were aligned and pairwise distance was computed using the Maximum Composite Likelihood model

(0.000; sequences had 100% nucleotide identity) to identify unique isolates (Tamura et al., 2004). A phylogenetic tree was created based on the partial 16S rRNA sequences of the unique isolates recovered and reference sequences. Reference sequences were obtained by targeting a known cultured isolate using the SeqMatch tool in RDP and retrieving the sequence from NCBI database. The final tree was constructed using the

Maximum Likelihood method based on the Tamura-Nei model (K. Tamura & Nei, 1993).

Bootstrap consensus tree inferred is from 1000 replicates.

Plant Growth Data Analysis and Statistics

Root to shoot ratio was obtained in 2020 only by dividing root biomass by aboveground biomass. The Shapiro-Wilks normality test was performed to examine normality of the different variables tested. Levene’s test was performed to determine homogeneity of variance. Log transformation was used on non-normal data. One-way analysis of variance (ANOVA) was used to determine significant differences in measurements between the treatments and blocks. Kruskal-Wallis test was performed on data that did not meet parametric assumptions after transformation. Significant F-values

(p<0.05) obtained led to using Fisher’s least square difference (LSD) test to determine

49 differences between treatments. R (R Core Team, 2014) was used to perform the statistical analyses and figures using “agricolae” (de Mendiburu, 2020), “car” (Fox &

Weisberg, 2019), “ggplot2” (H. Wickham, 2016), "lsmeans" (Lenth, 2016) and "readr"

(Wickham et al., 2018) packages.

Results

Effect of Soil Dilution Treatments on Aboveground Biomass and Height of Tomato Seedlings

Two greenhouse experiments were conducted in an aim to determine the impact of chicken-grazed soil dilutions on tomato seedlings and stem endophytes. In the 2018 experiment, there were significant differences in aboveground biomass and height of the seedlings of both varieties grown in the different soil dilutions (Table 2.2). The most diluted chicken-grazed soil treatments (100% sterile soil substrate, 25% and 50% chicken-grazed soil) had significantly higher aboveground biomass and height compared to the least diluted soil treatments (75% and 100% chicken-grazed soil) (Figure 2.4).

Effect of Soil Slurry Dilution Treatments on Aboveground Biomass, Root Biomass, Height and Number of Compound Leaves of Tomato Seedlings

In 2020, for each chicken-grazed soil slurry dilution preparation CFUs were quantified. For each dilution, the amount of CFUs decreases, showing how there is less quantifiable bacteria as it is closer to the control treatment, as well as a decreasing

50 concentration of soil bacteria that can establish on the growing substrate and transmit into the plant (Table 2.3). There was a significant treatment effect on tomato seedling height, number of compound leaves and root to shoot ratio (Table 2.4). According to the Fisher's

LSD test, the most diluted treatment (10-4) had a significantly higher height compared to the other treatments and the non-inoculated control (Table 2.5). Except for height, no significant differences were observed between the non-inoculated control and the soil slurry dilutions for all measurements across treatments. Overall, the highest aboveground biomass and height was observed on the most diluted treatment (10-4) and the lowest measurements were seen in the least diluted treatment (10-1). However, a significant block effect was observed on aboveground and root biomass. Measurements from tomato seedlings grown in block B3 were the highest.

Impact of Dilution Treatments on Cultured Endophytic Bacteria Composition

Endophytic bacteria were recovered from surface-sterilized stem of tomato seedlings from all treatments using two sources of nutrient media. In 2018, a total of 87 endophytic bacteria were recovered and identified at a genus level. By computing the pairwise distance of each 87 bacterial sequences, 35 of these represent unique taxa. The classification of these 35 taxa, based on >97% similarity to GenBank database entries queried through blastn is shown in Table 2.6. Recovered isolates belong to 15 genera and four phyla: Actinobacteria, Firmicutes, Proteobacteria and Bacteroidetes (Figure 2.5).

One isolate could only be identified to the family level as Enterobacteriaceae

51 consistently in NCBI and RDP. However, these differed at genus level. An unknown bacterium in NCBI was the closest match to one of the isolates, whereas RDP classified it as Enterobacter sp. Hence, this isolate will be referred as Enterobacteriaceae 1. The top three most abundant genera recovered were Bacillus, Microbacterium and Paenibacillus, representing 60%, 14% and 7% of isolates, respectively. Taxa belonging to Bacillus were found on all treatments (Figure 2.6). Many of the taxa were present in 1-3 treatments.

Bacterial endophytes assigned to the genera Ochrobactrum, Chryseobacterium and

Cellusimicrobium were only found in 100% sterile soil substrate.

A phylogenetic reconstruction using the unique taxa recovered and reference sequences was generated based on 16S rRNA gene sequence alignments (Figure 2.7).

Most of the taxa recovered were closely related to known bacteria associated to soil, rhizosphere and plant sources. However, some taxa, such as Bacillus 3, Paenibacillus 2,

Microbacterium 1-4, 1, and Enterobacteriaceae 1 were closely related to bacteria of clinical relevance. Some of these bacteria were only recovered from treatments containing chicken-grazed soils (25% through 100% chicken-grazed soil).

Lysinibacillus 1-2, Arthrobacter 1, Streptomyces 1, Pelomonas 1, Pseudomonas 1,

Ochrobactrum 1, Sphingobium 1-2, Chryseobacterium 1 and Flavobacterium 1-2 did not have closely related reference taxa that originated from clinical settings.

To differentiate within species of each recovered genera, the rpoB region was selected. Two primer pairs were used to target rpoB: rpoB-f/rpoB-r and Univ rpoB-

52

F/Univ rpoB-R. The former targets Bacillus spp. and the latter is a universal primer.

When the bacterial isolates were sequenced by targeting the rpoB gene region, 25 out of

87 amplified. From the sequences recovered from the remaining 25 isolates, 11 amplified using primer pair rpoB-f/rpoB-r and 14 amplified using primer pair Univ rpoB-F/Univ rpoB-R (Table 2.7). All isolates amplified using primer pair rpoB-f/rpoB-r showed a high similarity (>97%, blastn) to NCBI database entries. Most of these isolates belong to the

Bacillus 3 taxa. Only 9 isolates using primer pair Univ rpoB-F/Univ rpoB-R showed a high similarity (>97%, blastn) to GenBank database entries. The rest of the isolates showed lower similarities (91–95%) to the closest match in NCBI database. Overall, only

29% of the isolates recovered were amplified with rpoB, while 16SrRNA was successful with 100% of the isolates, and with the studied primers, rpoB did not provide greater taxonomic resolution of the tested isolates.

In 2020, a total of 263 endophytic bacterial isolates were recovered from surface sterilized stem tissue of tomatoes grown in substrate inoculated with different soil slurry dilutions. Isolates were stored at -80oC (Table 2.8). Future work will require the identification of these isolates.

Discussion

In our experiments, chicken-grazed soil introduced in the planting substrate impacted both tomato seedling vigor and bacterial endophytes. The greater percent of

53 chicken grazed soil, the greater the stunting of tomato seedlings. Similarly, a subset of isolates recovered in these experiments differed by treatment; however, most of the recovered isolates identified in 2018 consisted of two dominant genera, Bacillus and

Microbacterium, the former being the only genus present across all treatments. Most of the recovered isolates were found to be phylogenetically similar to formerly cultured plant endophytes and bacteria found in soil. However, some isolates were similar to bacteria recovered in human clinical settings. Microbial community selection is known to be influenced by plant genotype, favoring microorganisms that confer beneficial effects to the host (Wickham, 2016). However, our results suggest the importance of planting substrate over plant genotype.

In the 2018 experiment, there was a significant decrease in aboveground biomass and height of both tomato seedling varieties grown in the dilutions containing high percentages of soil from CGPPs. In 2018, soils from CGPPs was mixed with an autoclaved soil substrate. Changes in biomass in 2018 might be due to nutrient availability being enhanced in the autoclaved soil substrate. Autoclaving soil is known to alter physicochemical characteristics of soil, such as changes in soil organic matter structure, pH and nutrient availability (Alphei & Scheu, 1993; Berns et al., 2008;

Serrasolsas & Khanna, 1995; Wolf et al., 1989). Miransari et al. (2009). observed a significant increase of nutrient uptake of mycorrhizal wheat grown in soil sterilized by autoclaving, compared to non-autoclaved soil. Although most studies on autoclaved soils focus on its effect in microbial activity, a few do provide information on autoclaving

54 effect on plant growth. For example, Mahmood et. al. (2014) found that wheat seedlings grown in soils sterilized through autoclaving and γ-irradiation, produced higher root biomass, length, surface area, volume and number of tips compared to those grown in the unsterilized soil.

The modifications of the experiment conducted in 2020 were an effort to develop a controlled inoculation protocol of diluted chicken-grazed soil slurry. The soil slurry assisted in reducing the impact of available nutrients, soil’s physical and chemical characteristics, and abundance of microorganisms. Previous studies have been successful at incorporating soil slurry inoculations, which assist in harnessing the benefits of soil and its microbiome. In other studies, upon inoculation with a soil slurry, microbial shifts in the rhizosphere and subsequently, in the plant host were observed (Jack et al., 2019;

Panke-Buisse et al., 2015). Similarly, previous studies have added a suspension of an inoculum to germinated seeds, to control microorganisms being introduced during germination (Armanhi et al., 2018; Niu et al., 2017).

While modifications to the protocol were performed in 2020, similar results were observed regarding tomato biomass, compared to the 2018 experiment. The least diluted soil slurry treatments had significantly lower tomato plant growth. Other than stunting, the tomato seedlings did not present any other symptomatology in above- or belowground tissues. Low tomato vigor observed in both 2018 and 2020 experiments might be due to the presence of chicken manure in the studied soils and its effect on nutrient availability

55 and phytotoxicity. Past research has shown that application of chicken manure caused stunting of potato plants (Conn & Lazarovits, 1999). Conn and Lazarovits (1999) also observed the effect to be site and soil specific and considered effects in pH as a possible explanation for site to site variation in response to chicken manure application and stunting. Riegel and Noe (2000) also observed a decreased of cotton plant growth in chicken manure amended pots compared with nonamended pots. The authors for both papers suggested the use of lower amounts of chicken manure to reduce potential phytotoxic effects. Chicken manure is known to be a nutrient rich medium, containing significant quantities of micronutrients, P, K and N, in forms of ammonia, nitrate, and nitrite, that can lead to a reduction in seed germination or a decrease in emergence (Moe et al., 2017; Roy and Kashem, 2014).

For tomato plants, nutrients such as nitrogen (N), phosphorous (P), potassium (K), calcium (Ca), magnesium (Mg) and sulfur (S), are needed in large amounts for normal growth and reproduction. Other nutrients, such as manganese (Mn), copper (Cu), zinc

(Zn), chlorine (Cl), boron (B), iron (Fe), and molybdenum (Mo), are needed in smaller amounts (Sainju et al., 2003). Chicken manure is known to contain all 13 of the essential tomato plant nutrients (Chastain et al., 2001). Excess level of P is not as harmful to tomato plants, but it can reduce the availability of some micronutrients, such as Fe, Zn,

Mn, and Cu. This is achieved by decreasing the nutrient’s solubility in the soil and translocation within the plant (Ismail et al., 1996; Needham, 1973). Deficiency of Fe can cause plant growth stunting (Sainju et al., 2003). A study found phytotoxic effects of

56 chicken manure on Banksia spinulosa due to high phosphorus concentrations (Aryantha et al., 2000). Ravindran et al. (2017) found chicken manure to have phytotoxic properties due to a recorded germination index of less than 50% on commercial vegetable crops, including tomato. The authors used a germination index to quantify phytotoxicity on commercial vegetable crops. The germination index is generated using measurements related to the performance of crop seedlings, relative seed germination and relative root elongation during the assay period. Crops, including tomato, with a germination index of less than 50% indicated a phytotoxic medium, compared to crops with 50% and 80% values which indicate moderate phytotoxicity. Nonetheless, germination was not affected in experiments conducted in 2018 and 2020. Incorporation of chicken manure, however, has shown improvement on plant growth depending on the rate and form of incorporation. Dikinya and Mufwanzala (2010) found initial increases and subsequent decreases of the dry biomass of spinach with application of increasing rate of chicken manure. Demir et al. (2010) found that application of pelletized poultry manure improved tomato shoot growth and fruit yield and increased leaf N concentrations at the harvest stage. Some studies lack specific information on rate and form of incorporated chicken manure, although results show an increase in plant growth (Chellemi et al., 2013; Maru et al., 2015). Therefore, the management of chicken-grazed soils through the protocols for

2018 and 2020 might impact the ideal amount of nutrient levels for tomato seedlings. We hypothesize that phytotoxic effects could be driving stunting in tomato seedlings grown in these experiments. However, for these experiments, nutrient availability was not measured.

57

Other than phytotoxicity, physical and chemical factors related to soil can also impact stunting. Soil compaction can affect root system architecture by limiting expansion and exploration of soil, ultimately affecting shoot growth (Lipiec & Hatano,

2003). Tomato plants grown in columns containing higher soil bulk density had reduced root surface area, total root volume and total root length (Tracy et al., 2012). Reduced crop growth due to compaction can be attributed to reduced water infiltration and crop water use efficiency (Radford et al., 2001). However, hypothesizing that soil compaction affected stunting of tomato seedlings would only apply to the 2018 experiment. Tomato seedlings from the 2020 experiment grew in the same substrate. Moreover, root to shoot ratio had a higher trend in treatments with highest dilutions. Plants are known to allocate its resources to aboveground compartments when growing in nutrient-rich environments due to competition for light. On the other hand, resources are allocated to belowground compartments when growing in nutrient-poor environments (Mašková & Herben, 2018).

Higher root to shoot ratios for tomato seedlings inoculated with the higher dilutions might reflect a nutrient-poor planting substrate.

An imbalance of plant hormones can impact stunting in plants. Abscisic acid

(ABA) production is related to stunting in plants, especially under water stress.

Regarding the 2018 experiment, there might be a reduced water infiltration in soil treatments with higher amounts of chicken-grazed soil, leading to water stress and eventually, increased amount of ABA in the tomato seedlings. ABA production increases during water stress, preventing excess ethylene production and subsequently, inhibiting

58 shoot growth (Sharp & LeNoble, 2002). Additionally, ABA induces stomatal closure and reduces water loss (Herrera-Medina et al., 2007). Increased concentrations of ABA are known to stunt growth of Brassica, maize and French beans (Cramer & Quarrie, 2002;

He & Cramer, 1996; Montero et al., 1998). There are microbes, such as Azospirillum spp. and Bacillus amyloliquefaciens RWL-1, involved in the production of ABA when plants are under water stress (Cohen et al., 2009; Shahzad et al., 2017). A smaller number of different taxa were recovered from plants grown in 100% chicken-grazed soil compared to other treatments. Bacillus constituted the majority of the endophytes recovered from this treatment. These microbes might have been facilitating in ABA production, causing stunting in growth.

Changes in soil physical and chemical factors, such as the addition of organic amendments, have an impact on endophytic communities (J. Hallmann et al., 1997). In

2018, a total of 87 endophytic bacteria belonging to 15 genera and four phyla

(Actinobacteria, Proteobacteria, Bacteroidetes and Firmicutes) were recovered from our experiment. Predominant plant endophytes typically belong to these major phyla

(Rosenblueth & Martínez-Romero, 2006). The origin of the isolates recovered from the stem, whether it is the soil or the tomato seed, is unknown. 60% of the bacterial endophytes were Bacillus and it was the only genus found across all treatments. Bacillus is as ubiquitous as it is studied for its beneficial modes of action, such as antagonism towards plant pathogens and pests and its plant growth promoting activity (Bai et al.,

2002; Hu et al., 2010; L. Ma et al., 2013; Melnick et al., 2008; Raymond et al., 2010).

59

Correspondingly, beneficial endophytic Bacillus have also been recovered from tomato plant tissues (Chauhan et al., 2014; Kefi et al., 2015; Thomas & Upreti, 2016).

Most of the endophytes recovered were closely related to known bacteria associated to soil, rhizosphere, and plant sources. Lysinibacillus 1-2, Arthrobacter 1,

Streptomyces 1, Pelomonas 1, Pseudomonas 1, Ochrobactrum 1, Sphingobium 1-2,

Chryseobacterium 1 and Flavobacterium 1-2 did not have reference 16S rDNA sequences that originated from clinical settings. Bacteria from some of these genera have already been identified as endophytes or rhizospheric microbes with beneficial properties towards plant hosts, such as tomato, and disease suppression (Abbamondi et al., 2016; Compant et al., 2011; Kunova et al., 2016; Palmieri et al., 2017). Isolates recovered from seedlings grown in 100% sterile soil substrate might be originally from the tomato seed.

Additionally, the only treatment where Bacillus was not as predominant was the 100% sterile soil substrate. Bacterial endophytes assigned to the genera Ochrobactrum,

Chryseobacterium and Cellusimicrobium were only found in 100% sterile soil substrate.

Although not reported to be common in tomato plants, these genera have been recovered from the rhizoplane to the phyllosphere of various crops (Akbaba & Ozaktan, 2018;

Akinsanya et al., 2015; Hardoim et al., 2012; Jeong et al., 2017; Lebuhn et al., 2000; Shi et al., 2010).

Based on a 16S rDNA sequence phylogeny, a subset of isolates were closely related to bacteria of human clinical relevance. These taxa, Bacillus 3, Paenibacillus 2,

60

Microbacterium 1-4, Cellulosimicrobium 1 and Enterobacteriaceae 1 also belong to genera with known species as plant endophytes as well as human pathogens (Beesley et al., 2010; Glaeser et al., 2013; Pablo Rodrigo Hardoim et al., 2013; McNeil et al., 2004;

Park et al., 2006; Yoon et al., 2007; Zhang et al., 2017). Various studies on the effect of chicken manure focused on the presence of human and animal pathogenic bacteria. Yang et. al has conducted studies on the presence of multiple antibiotic-resistant bacteria in chicken manure and manure fertilized vegetables (Qingxiang Yang et al., 2016; Yang et al., 2014). The implementation of chicken manure, as a source of nutrients and pathogenic microorganisms, can also pose environmental concerns due to runoff from agricultural land (Sistani et al., 2009).

One isolate recovered from a tomato seedling grown in 50% chicken-grazed soil treatment could only be identified to the family level as Enterobacteriaceae. While this family is predominantly known to contain taxa associated to human pathogens, other taxa in Enterobacteriaceae have been identified as plant endophytes (Bulgari et al., 2009; de

Abreu et al., 2017; Magnani et al., 2010). Bacteria in the family Enterobacteriaceae have shown a wide functional diversity, by expressing traits such as auxin production, phosphate solubilization, siderophore production and nitrogen-fixation (Khalaf &

Raizada, 2016; T. Luo et al., 2016; Zimmer et al., 1994). Enterobacteriaceae has also been found to be part of the endophytic community of tomato plants (Romero et al.,

2014). Additionally, studies have been conducted attempting to understand colonization patterns of bacteria from this family (Leite et al., 2013; Lopez-Fernandez et al., 2016).

61

Although some studies indicate beneficial roles of bacteria in this family, there is concern surrounding the presence of these bacteria and their ability to cause infectious diseases in animals and humans. For example, bacteria from the family Enterobacteriaceae were isolated from various wild and cultivated plants, including tomato, and were shown to have a broad spectrum of resistance to antibiotics and were highly adhesive to human erythrocytes (Markova et al., 2005).

Considering that certain endophytic bacteria are closely related to known human and animal pathogens, it is of importance to identify species and strain level differences.

Although the identification of 16S rRNA gene region allows for successful discrimination of diverse microbial taxa, it has shown difficult to assign species level (Kim et al., 2011). The use of other marker genes, such as rpoB (encodes the

RNA polymerase β Subunit) or gyrB (encodes the subunit B protein of DNA gyrase), can better assist in defining phylogenetic relationships by (Arfaoui et al., 2019; Yamamoto &

Harayama, 1995). In an attempt to determine the recovered endophytic bacteria’s species, the rpoB region was selected to be amplified due to it exhibiting higher levels of divergence compared to 16S rRNA gene region (Mollet et al., 1997; Vos et al., 2012).

Two primer pairs were used where one targets Bacillus spp. and the other is a universal primer. When the bacterial isolates were sequenced by targeting the rpoB gene region, 25 out of 87 amplified. The primers that target the genus Bacillus were not as specific as expected and the universal primers were not as broad considering the low number of recovered isolates with an amplification product for the rpoB gene. Alternatively, other

62 universally conserved genes can be used to obtain genetic information from the recovered isolates (Santos & Ochman, 2004). Methods, such as multi locus sequence analysis, can also assist in understanding phylogenetic relationships of species within a genus or genera within a family, especially in cases where the genus has species that are endophytes and human and animal pathogens (Brady et al., 2008; Glaeser & Kämpfer,

2015; Gonçalves et al., 2018).

The bacterial endophytes recovered in this study can be further studied to uncover their potential as plant growth promoting agents or antagonists of plant pathogens.

However, further work must be done in first differentiating possible plant and human pathogens. Changes in management of chicken-grazed soil could potentially assist in controlling bacteria detrimental to human and plant health. Composting chicken manure, for example, lead to a decrease in potential pathogenic bacteria, antibiotics and antibiotic resistance genes (Deng et al., 2020). With the proper protocols in place, the effectiveness of chicken grazing on subsequent plant growth and yield, as well as endophyte diversity and composition, can be determined. Our findings also indicate the need to assess parameters such as rate and form of incorporation of chicken manure for success in managing soil fertility and minimizing phytotoxicity. Experiments with controlled inoculation, as well as controlled incorporation, of chicken-grazed soils are helpful in interpreting these parameters and their impact on plant growth, yield, and its associated microbiome. However, it is important to determine soil factors such as nutrient availability as well as chemical and physical characteristics to truly create reproducible

63 conditions. Consequently, these conditions can assist in obtaining causal links between bacterial endophytes and their potential impact on plant health.

References

Abbamondi, G. R., Tommonaro, G., Weyens, N., Thijs, S., Sillen, W., Gkorezis, P., Iodice, C., de Melo Rangel, W., Nicolaus, B., & Vangronsveld, J. (2016). Plant growth-promoting effects of rhizospheric and endophytic bacteria associated with different tomato cultivars and new tomato hybrids. Chemical and Biological Technologies in Agriculture, 3(1), 1. https://doi.org/10.1186/s40538- 015-0051-3

Acosta-Martínez, V., Bell, C. W., Morris, B. E. L., Zak, J., & Allen, V. G. (2010). Long-term soil microbial community and enzyme activity responses to an integrated cropping-livestock system in a semi-arid region☆. Agriculture, Ecosystems & Environment, 137(3–4), 231–240. https://doi.org/10.1016/j.agee.2010.02.008

Akbaba, M., & Ozaktan, H. (2018). Biocontrol of angular leaf spot disease and colonization of cucumber (Cucumis sativus L.) by endophytic bacteria. Egyptian Journal of Biological Pest Control, 28(1), 14. https://doi.org/10.1186/s41938- 017-0020-1

Akinsanya, M. A., Goh, J. K., Lim, S. P., & Ting, A. S. Y. (2015). Diversity, antimicrobial and antioxidant activities of culturable bacterial endophyte communities in Aloe vera. FEMS Microbiology Letters, 362(23). https://doi.org/10.1093/femsle/fnv184

Alphei, J., & Scheu, S. (1993). Effects of biocidal treatments on biological and nutritional properties of a mull-structured woodland soil. In L. Brussaard & M. J. Kooistra (Eds.), Soil Structure/Soil Biota Interrelationships (pp. 435–448). Elsevier. https://doi.org/10.1016/B978-0-444-81490-6.50035-2

Altschul, S. F., Gish, W., Miller, W., Myers, E. W., & Lipman, D. J. (1990). Basic local alignment search tool. Journal of Molecular Biology, 215(3), 403–410. https://doi.org/10.1016/S0022-2836(05)80360-2

Arfaoui, M., Vallance, J., Bruez, E., Rezgui, A., Melki, I., Chebil, S., Sadfi-Zouaoui, N., & Rey, P. (2019). Isolation, identification and in vitro characterization of grapevine rhizobacteria to control ochratoxigenic Aspergillus spp. On grapes.

64

Biological Control, 129, 201–211. https://doi.org/10.1016/j.biocontrol.2018.10.019

Armanhi, J. S. L., de Souza, R. S. C., Damasceno, N. de B., de Araújo, L. M., Imperial, J., & Arruda, P. (2018). A Community-Based Culture Collection for Targeting Novel Plant Growth-Promoting Bacteria from the Sugarcane Microbiome. Frontiers in Plant Science, 8. https://doi.org/10.3389/fpls.2017.02191

Aryantha, I. P., Cross, R., & Guest, D. I. (2000). Suppression of Phytophthora cinnamomi in Potting Mixes Amended with Uncomposted and Composted Animal Manures. Phytopathology®, 90(7), 775–782. https://doi.org/10.1094/PHYTO.2000.90.7.775

Bacon, C. W., & White, J. (2000). Microbial Endophytes. CRC Press.

Bai, Y., D’Aoust, F., Smith, D. L., & Driscoll, B. T. (2002). Isolation of plant-growth- promoting Bacillus strains from soybean root nodules. Canadian Journal of Microbiology, 48(3), 230–238. https://doi.org/10.1139/w02-014

Beesley, C. A., Vanner, C. L., Helsel, L. O., Gee, J. E., & Hoffmaster, A. R. (2010). Identification and characterization of clinical Bacillus spp. Isolates phenotypically similar to Bacillus anthracis. FEMS Microbiology Letters, 313(1), 47–53. https://doi.org/10.1111/j.1574-6968.2010.02120.x

Bergougnoux, V. (2014). The history of tomato: From domestication to biopharming. Biotechnology Advances, 32(1), 170–189. https://doi.org/10.1016/j.biotechadv.2013.11.003

Berns, A. E., Philipp, H., Narres, H.-D., Burauel, P., Vereecken, H., & Tappe, W. (2008). Effect of gamma-sterilization and autoclaving on soil organic matter structure as studied by solid state NMR, UV and fluorescence spectroscopy. European Journal of Soil Science, 59(3), 540–550. https://doi.org/10.1111/j.1365-2389.2008.01016.x

Bodenhausen, N., Bortfeld-Miller, M., Ackermann, M., & Vorholt, J. A. (2014). A Synthetic Community Approach Reveals Plant Genotypes Affecting the Phyllosphere Microbiota. PLoS Genetics, 10(4). https://doi.org/10.1371/journal.pgen.1004283

Brady, C., Cleenwerck, I., Venter, S., Vancanneyt, M., Swings, J., & Coutinho, T. (2008). Phylogeny and identification of Pantoea species associated with plants, humans and the natural environment based on multilocus sequence analysis (MLSA). Systematic and Applied Microbiology, 31(6), 447–460. https://doi.org/10.1016/j.syapm.2008.09.004 65

Bulgari, D., Casati, P., Brusetti, L., Quaglino, F., Brasca, M., Daffonchio, D., & Bianco, P. A. (2009). Endophytic bacterial diversity in grapevine (Vitis vinifera L.) leaves described by 16S rRNA gene sequence analysis and length heterogeneity- PCR. The Journal of Microbiology, 47(4), 393–401. https://doi.org/10.1007/s12275-009-0082-1

Case, R. J., Boucher, Y., Dahllöf, I., Holmström, C., Doolittle, W. F., & Kjelleberg, S. (2007). Use of 16S rRNA and rpoB genes as molecular markers for microbial ecology studies. Applied and Environmental Microbiology, 73(1), 278–288. https://doi.org/10.1128/AEM.01177-06

Chastain, J. P., Camberato, J. J., & Skewes, P. (2001). Poultry Manure Production and Nutrient Content. 17.

Chauhan, A., Guleria, S., Walia, A., Mahajan, R., Verma, S., & Shirkot, C. K. (2014). Isolation and characterization of Bacillus sp. With their effect on growth of tomato seedlings. Indian Journal of Agricultural Biochemistry, 27(2), 193–201.

Chelius, M. K., & Triplett, E. W. (2000). Immunolocalization of Dinitrogenase Reductase Produced by Klebsiella pneumoniae in Association with Zea mays L. Applied and Environmental Microbiology, 66(2), 783–787. https://doi.org/10.1128/AEM.66.2.783-787.2000

Chellemi, D. O., Rosskopf, E. N., & Kokalis-Burelle, N. (2013). The Effect of Transitional Organic Production Practices on Soilborne Pests of Tomato in a Simulated Microplot Study. Phytopathology, 103(8), 792–801. https://doi.org/10.1094/PHYTO-09-12-0243-R

Chen, C., Chen, H. Y. H., Chen, X., & Huang, Z. (2019). Meta-analysis shows positive effects of plant diversity on microbial biomass and respiration. Nature Communications, 10(1), 1–10. https://doi.org/10.1038/s41467-019-09258-y

Chihaoui, S.-A., Trabelsi, D., Jdey, A., Mhadhbi, H., & Mhamdi, R. (2015). Inoculation of Phaseolus vulgaris with the nodule-endophyte Agrobacterium sp. 10C2 affects richness and structure of rhizosphere bacterial communities and enhances nodulation and growth. Archives of Microbiology, 197(6), 805–813. https://doi.org/10.1007/s00203-015-1118-z

Cohen, A. C., Travaglia, C. N., Bottini, R., & Piccoli, P. N. (2009). Participation of abscisic acid and gibberellins produced by endophytic Azospirillum in the alleviation of drought effects in maize. Botany, 87(5), 455–462. https://doi.org/10.1139/B09-023

66

Compant, S., Clément, C., & Sessitsch, A. (2010). Plant growth-promoting bacteria in the rhizo- and endosphere of plants: Their role, colonization, mechanisms involved and prospects for utilization. Soil Biology and Biochemistry, 42(5), 669–678. https://doi.org/10.1016/j.soilbio.2009.11.024

Compant, S., Mitter, B., Colli-Mull, J. G., Gangl, H., & Sessitsch, A. (2011). Endophytes of Grapevine Flowers, Berries, and Seeds: Identification of Cultivable Bacteria, Comparison with Other Plant Parts, and Visualization of Niches of Colonization. Microbial Ecology, 62(1), 188–197. https://doi.org/10.1007/s00248-011-9883-y

Conn, K. L., & Lazarovits, G. (1999). Impact of animal manures on verticillium wilt, potato scab, and soil microbial populations. Canadian Journal of Plant Pathology, 21(1), 81–92. https://doi.org/10.1080/07060661.1999.10600089

Connon, S. A., & Giovannoni, S. J. (2002). High-Throughput Methods for Culturing Microorganisms in Very-Low-Nutrient Media Yield Diverse New Marine Isolates. Applied and Environmental Microbiology, 68(8), 3878–3885. https://doi.org/10.1128/AEM.68.8.3878-3885.2002

Cordero, J., de Freitas, J. R., & Germida, J. J. (2019). Bacterial microbiome associated with the rhizosphere and root interior of crops in Saskatchewan, Canada. Canadian Journal of Microbiology, 66(1), 71–85. https://doi.org/10.1139/cjm- 2019-0330

Cramer, G. R., & Quarrie, S. A. (2002). Abscisic acid is correlated with the leaf growth inhibition of four genotypes of maize differing in their response to salinity. Functional Plant Biology, 29(1), 111–115. https://doi.org/10.1071/pp01131 de Abreu, C. S., Figueiredo, J. E. F., Oliveira, C. A., dos Santos, V. L., Gomes, E. A., Ribeiro, V. P., Barros, B. A., Lana, U. G. P., & Marriel, I. E. (2017). Maize endophytic bacteria as mineral phosphate solubilizers. Genetics and Molecular Research, 16(1). https://doi.org/10.4238/gmr16019294

De Clerck, E., & De Vos, P. (2004). Genotypic diversity among Bacillus licheniformis strains from various sources. FEMS Microbiology Letters, 231(1), 91–98. https://doi.org/10.1016/S0378-1097(03)00935-2 de Faccio Carvalho, P. C., Anghinoni, I., de Moraes, A., de Souza, E. D., Sulc, R. M., Lang, C. R., Flores, J. P. C., Terra Lopes, M. L., da Silva, J. L. S., Conte, O., de Lima Wesp, C., Levien, R., Fontaneli, R. S., & Bayer, C. (2010). Managing grazing animals to achieve nutrient cycling and soil improvement in no-till integrated systems. Nutrient Cycling in Agroecosystems, 88(2), 259–273. https://doi.org/10.1007/s10705-010-9360-x 67

de Mendiburu, F. (2020). agricolae: Statistical Procedures for Agricultural Research. (Version R package version 1.3-2) [Computer software]. https://CRAN.R- project.org/package=agricolae

Demir, K., Sahin, O., Kadioglu, Y. K., Pilbeam, D. J., & Gunes, A. (2010). Essential and non-essential element composition of tomato plants fertilized with poultry manure. Scientia Horticulturae, 127(1), 16–22. https://doi.org/10.1016/j.scienta.2010.08.009

Deng, W., Zhang, A., Chen, S., He, X., Jin, L., Yu, X., Yang, S., Li, B., Fan, L., Ji, L., Pan, X., & Zou, L. (2020). Heavy metals, antibiotics and nutrients affect the bacterial community and resistance genes in chicken manure composting and fertilized soil. Journal of Environmental Management, 257, 109980. https://doi.org/10.1016/j.jenvman.2019.109980

Deng, Z.-S., Zhang, B.-C., Qi, X.-Y., Sun, Z.-H., He, X.-L., Liu, Y.-Z., Li, J., Chen, K.- K., & Lin, Z.-X. (2019). Root-Associated Endophytic Bacterial Community Composition of Pennisetum sinese from Four Representative Provinces in China. Microorganisms, 7(2). https://doi.org/10.3390/microorganisms7020047

Devendra, C., & Thomas, D. (2002). Crop–animal interactions in mixed farming systems in Asia. Agricultural Systems, 71(1–2), 27–40. https://doi.org/10.1016/S0308-521X(01)00034-8

Dikinya, O., & Mufwanzala, N. (2010). Chicken manure-enhanced soil fertility and productivity: Effects of application rates. Journal of Soil Science and Environmental Management, 1(3), 46–54.

Ding, T., Palmer, M. W., & Melcher, U. (2013). Community terminal restriction fragment length polymorphisms reveal insights into the diversity and dynamics of leaf endophytic bacteria. BMC Microbiology, 13(1), 1. https://doi.org/10.1186/1471-2180-13-1

Fox, J., & Weisberg, S. (2019). An {R} Companion to Applied Regression (Third). Sage. https://socialsciences.mcmaster.ca/jfox/Books/Companion/

Gardner, J. C., Faulkner, D. B., & Hargrove, W. L. (1991). Use of cover crops with integrated crop-livestock production systems. Cover Crops for Clean Water. Soil and Water Conserv. Soc., Ankeny, IA, 185–191.

Gardner, J. M., Feldman, A. W., & Zablotowicz, R. M. (1982). Identity and behavior of xylem-residing bacteria in rough lemon roots of Florida citrus trees. Applied and Environmental Microbiology, 43(6), 1335–1342. 68

Gdanetz, K., & Trail, F. (2017). The Wheat Microbiome Under Four Management Strategies, and Potential for Endophytes in Disease Protection. Phytobiomes Journal, 1(3), 158–168. https://doi.org/10.1094/PBIOMES-05-17-0023-R

Genomics Workbench version 20.0. (2020). https://digitalinsights.qiagen.com

Gerberastraat, B. V., & Aalsmeer, R. A. (2001). 4Peaks. Nucleobytes. https://nucleobytes.com/4peaks/index.html

Glaeser, S. P., Falsen, E., Busse, H.-J., & Kämpfer, P. (2013). Paenibacillus vulneris sp. Nov., isolated from a necrotic wound. International Journal of Systematic and Evolutionary Microbiology, 63(Pt 2), 777–782. https://doi.org/10.1099/ijs.0.041210-0

Glaeser, S. P., & Kämpfer, P. (2015). Multilocus sequence analysis (MLSA) in prokaryotic taxonomy. Systematic and Applied Microbiology, 38(4), 237–245. https://doi.org/10.1016/j.syapm.2015.03.007

Gonçalves, R., Balbi-Peña, M., Soman, J., Maringoni, A., Taghouti, G., Fischer-Le Saux, M., & Portier, P. (2018). Genetic diversity of Curtobacterium flaccumfaciens revealed by multilocus sequence analysis. European Journal of Plant Pathology. https://doi.org/10.1007/s10658-018-01648-0

Gupta, V., Rai, P. K., & Risam, K. S. (2012). Integrated crop-livestock farming systems: A strategy for resource conservation and environmental sustainability. Indian Research Journal of Extension Education, Special Issue, 2, 49–54.

Hallmann, J., Quadt-Hallmann, A., Mahaffee, W. F., & Kloepper, J. W. (1997). Bacterial endophytes in agricultural crops. Canadian Journal of Microbiology, 43(10), 895–914. https://doi.org/10.1139/m97-131

Hallmann, Johannes, Berg, G., & Schulz, B. (2006). Isolation Procedures for Endophytic Microorganisms. In B. J. E. Schulz, C. J. C. Boyle, & T. N. Sieber (Eds.), Microbial Root Endophytes (pp. 299–319). Springer. https://doi.org/10.1007/3-540-33526-9_17

Hardoim, Pablo R., Hardoim, C. C. P., Overbeek, L. S. van, & Elsas, J. D. van. (2012). Dynamics of Seed-Borne Rice Endophytes on Early Plant Growth Stages. PLOS ONE, 7(2), e30438. https://doi.org/10.1371/journal.pone.0030438

Hardoim, Pablo R., van Overbeek, L. S., Berg, G., Pirttilä, A. M., Compant, S., Campisano, A., Döring, M., & Sessitsch, A. (2015). The Hidden World within Plants: Ecological and Evolutionary Considerations for Defining Functioning of 69

Microbial Endophytes. Microbiology and Molecular Biology Reviews, 79(3), 293–320. https://doi.org/10.1128/MMBR.00050-14

Hardoim, Pablo Rodrigo, Nazir, R., Sessitsch, A., Elhottová, D., Korenblum, E., van Overbeek, L. S., & van Elsas, J. D. (2013). The new species Enterobacter oryziphilus sp. Nov. And Enterobacter oryzendophyticus sp. Nov. Are key inhabitants of the endosphere of rice. BMC Microbiology, 13, 164. https://doi.org/10.1186/1471-2180-13-164

He, T., & Cramer, G. R. (1996). Abscisic acid concentrations are correlated with leaf area reductions in two salt-stressed rapid-cycling Brassica species. Plant and Soil, 179(1), 25–33. https://doi.org/10.1007/BF00011639

Herrera-Medina, M. J., Steinkellner, S., Vierheilig, H., Ocampo Bote, J. A., & García Garrido, J. M. (2007). Abscisic acid determines arbuscule development and functionality in the tomato arbuscular mycorrhiza. New Phytologist, 175(3), 554–564. https://doi.org/10.1111/j.1469-8137.2007.02107.x

Hoffmann, H., & Roggenkamp, A. (2003). Population Genetics of the Nomenspecies Enterobacter cloacae. Appl. Environ. Microbiol., 69(9), 5306–5318. https://doi.org/10.1128/AEM.69.9.5306-5318.2003

Hu, H. Q., Li, X. S., & He, H. (2010). Characterization of an antimicrobial material from a newly isolated Bacillus amyloliquefaciens from mangrove for biocontrol of Capsicum bacterial wilt. Biological Control, 54(3), 359–365. https://doi.org/10.1016/j.biocontrol.2010.06.015

Ismail, A.-S., Eissa, A., El-Beltagy, A., & Abou Hadid, A. (1996). Tomato growth in calcareous soils in relation to forms and levels of some macro- and micronutrients. Acta Horticulturae, 434, 85–94. https://doi.org/10.17660/ActaHortic.1996.434.9

Jack, C. N., Wozniak, K. J., Porter, S. S., & Friesen, M. L. (2019). Rhizobia protect their legume hosts against soil-borne microbial antagonists in a host-genotype- dependent manner. Rhizosphere, 9, 47–55. https://doi.org/10.1016/j.rhisph.2018.11.005

Jeong, J.-J., Lee, D. W., Park, B., Sang, M. K., Choi, I.-G., & Kim, K. D. (2017). Chryseobacterium cucumeris sp. Nov., an endophyte isolated from cucumber (Cucumis sativus L.) root, and emended description of Chryseobacterium arthrosphaerae. International Journal of Systematic and Evolutionary Microbiology, 67(3), 610–616. https://doi.org/10.1099/ijsem.0.001670

70

Jiao, N. (2013). Developing a Novel Approach of rpoB Gene as a Powerful Biomarker for the Environmental Microbial Diversity. Geomicrobiology Journal, 30(2), 108–119. https://doi.org/10.1080/01490451.2011.653090

Kawasaki, A., Donn, S., Ryan, P. R., Mathesius, U., Devilla, R., Jones, A., & Watt, M. (2016). Microbiome and Exudates of the Root and Rhizosphere of Brachypodium distachyon, a Model for Wheat. PloS One, 11(10), e0164533. https://doi.org/10.1371/journal.pone.0164533

Kefi, A., Slimene, I. B., Karkouch, I., Rihouey, C., Azaeiz, S., Bejaoui, M., Belaid, R., Cosette, P., Jouenne, T., & Limam, F. (2015). Characterization of endophytic Bacillus strains from tomato plants (Lycopersicon esculentum) displaying antifungal activity against Botrytis cinerea Pers. World Journal of Microbiology and Biotechnology, 31(12), 1967–1976. https://doi.org/10.1007/s11274-015- 1943-x

Khalaf, E. M., & Raizada, M. N. (2016). Taxonomic and functional diversity of cultured seed associated microbes of the cucurbit family. BMC Microbiology, 16(1), 131. https://doi.org/10.1186/s12866-016-0743-2

Kim, M., Morrison, M., & Yu, Z. (2011). Evaluation of different partial 16S rRNA gene sequence regions for phylogenetic analysis of microbiomes. Journal of Microbiological Methods, 84(1), 81–87. https://doi.org/10.1016/j.mimet.2010.10.020

Kumar, S., Stecher, G., & Tamura, K. (2016). MEGA7: Molecular Evolutionary Genetics Analysis Version 7.0 for Bigger Datasets. Molecular Biology and Evolution, 33(7), 1870–1874. https://doi.org/10.1093/molbev/msw054

Kunova, A., Bonaldi, M., Saracchi, M., Pizzatti, C., Chen, X., & Cortesi, P. (2016). Selection of Streptomyces against soil borne fungal pathogens by a standardized dual culture assay and evaluation of their effects on seed germination and plant growth. BMC Microbiology, 16(1), 272. https://doi.org/10.1186/s12866-016- 0886-1

Lebuhn, M., Achouak, W., Schloter, M., Berge, O., Meier, H., Barakat, M., Hartmann, A., & Heulin, T. (2000). Taxonomic characterization of Ochrobactrum sp. Isolates from soil samples and wheat roots, and description of Ochrobactrum tritici sp. Nov. And Ochrobactrum grignonense sp. Nov. International Journal of Systematic and Evolutionary Microbiology, 50(6), 2207–2223. https://doi.org/10.1099/00207713-50-6-2207

Leite, H. A. C., Silva, A. B., Gomes, F. P., Gramacho, K. P., Faria, J. C., de Souza, J. T., & Loguercio, L. L. (2013). Bacillus subtilis and Enterobacter cloacae 71

endophytes from healthy Theobroma cacao L. trees can systemically colonize seedlings and promote growth. Applied Microbiology and Biotechnology, 97(6), 2639–2651. https://doi.org/10.1007/s00253-012-4574-2

Lenth, R. V. (2016). Least-Squares Means: The R Package lsmeans. Journal of Statistical Software, 69(1), 1–33. https://doi.org/10.18637/jss.v069.i01

Lipiec, J., & Hatano, R. (2003). Quantification of compaction effects on soil physical properties and crop growth. Geoderma, 116(1), 107–136. https://doi.org/10.1016/S0016-7061(03)00097-1

Liu, L., Chen, H., Cai, P., Liang, W., & Huang, Q. (2009). Immobilization and phytotoxicity of Cd in contaminated soil amended with chicken manure compost. Journal of Hazardous Materials, 163(2–3), 563–567. https://doi.org/10.1016/j.jhazmat.2008.07.004

Lopez-Fernandez, S., Compant, S., Vrhovsek, U., Bianchedi, P. L., Sessitsch, A., Pertot, I., & Campisano, A. (2016). Grapevine colonization by endophytic bacteria shifts secondary metabolism and suggests activation of defense pathways. Plant and Soil, 405(1), 155–175. https://doi.org/10.1007/s11104-015-2631-1

Luo, J., Wyatt, J., van der Weerden, T. J., Thomas, S. M., de Klein, C. A. M., Li, Y., Rollo, M., Lindsey, S., Ledgard, S. F., Li, J., Ding, W., Qin, S., Zhang, N., Bolan, N., Kirkham, M. B., Bai, Z., Ma, L., Zhang, X., Wang, H., … Rys, G. (2017). Chapter Five—Potential Hotspot Areas of Nitrous Oxide Emissions From Grazed Pastoral Dairy Farm Systems. In D. L. Sparks (Ed.), Advances in Agronomy (Vol. 145, pp. 205–268). Academic Press. https://doi.org/10.1016/bs.agron.2017.05.006

Luo, T., Ou‐Yang, X.-Q., Yang, L.-T., Li, Y.-R., Song, X.-P., Zhang, G.-M., Gao, Y.-J., Duan, W.-X., & An, Q. (2016). Raoultella sp. Strain L03 fixes N2 in association with micropropagated sugarcane plants. Journal of Basic Microbiology, 56(8), 934–940. https://doi.org/10.1002/jobm.201500738

Ma, L., Cao, Y. H., Cheng, M. H., Huang, Y., Mo, M. H., Wang, Y., Yang, J. Z., & Yang, F. X. (2013). Phylogenetic diversity of bacterial endophytes of Panax notoginseng with antagonistic characteristics towards pathogens of root-rot disease complex. Antonie van Leeuwenhoek, 103(2), 299–312. https://doi.org/10.1007/s10482-012-9810-3

Ma, Y., Rajkumar, M., Luo, Y., & Freitas, H. (2011). Inoculation of endophytic bacteria on host and non-host plants—Effects on plant growth and Ni uptake. Journal of Hazardous Materials, 195, 230–237. https://doi.org/10.1016/j.jhazmat.2011.08.034 72

Ma, Y., Rajkumar, M., Zhang, C., & Freitas, H. (2016). Inoculation of Brassica oxyrrhina with plant growth promoting bacteria for the improvement of heavy metal phytoremediation under drought conditions. Journal of Hazardous Materials, 320, 36–44. https://doi.org/10.1016/j.jhazmat.2016.08.009

Magnani, G. S., Didonet, C. M., Cruz, L. M., Picheth, C. F., Pedrosa, F. O., & Souza, E. M. (2010). Diversity of endophytic bacteria in Brazilian sugarcane. Genetics and Molecular Research, 9(1), 250–258. https://doi.org/10.4238/vol9-1gmr703

Mahmood, T., Mehnaz, S., Fleischmann, F., Ali, R., Hashmi, Z. H., & Iqbal, Z. (2014). Soil sterilization effects on root growth and formation of rhizosheaths in wheat seedlings. Pedobiologia, 57(3), 123–130. https://doi.org/10.1016/j.pedobi.2013.12.005

Markova, Yu. A., Romanenko, A. S., & Dukhanina, A. V. (2005). Isolation of Bacteria of the Family Enterobacteriaceae from Plant Tissues. Microbiology, 74(5), 575– 578. https://doi.org/10.1007/s11021-005-0105-9

Maru, A., Haruna, O. A., & Charles Primus, W. (2015). Coapplication of Chicken Litter Biochar and Urea Only to Improve Nutrients Use Efficiency and Yield of Oryza sativa L. Cultivation on a Tropical Acid Soil [Research Article]. The Scientific World Journal; Hindawi. https://doi.org/10.1155/2015/943853

Mašková, T., & Herben, T. (2018). Root:shoot ratio in developing seedlings: How seedlings change their allocation in response to seed mass and ambient nutrient supply. Ecology and Evolution, 8(14), 7143–7150. https://doi.org/10.1002/ece3.4238

McGilloway, D. A. (2005). Evolution of integrated crop-livestock production systems. In Grassland: A global resource: A Global Resource (pp. 137–148). Wageningen Academic Publishers.

McNeil, M. M., Brown, J. M., Carvalho, M. E., Hollis, D. G., Morey, R. E., & Reller, L. B. (2004). Molecular epidemiologic evaluation of endocarditis due to Oerskovia turbata and CDC group A-3 associated with contaminated homograft valves. Journal of Clinical Microbiology, 42(6), 2495–2500. https://doi.org/10.1128/JCM.42.6.2495-2500.2004

Melnick, R. L., Zidack, N. K., Bailey, B. A., Maximova, S. N., Guiltinan, M., & Backman, P. A. (2008). Bacterial endophytes: Bacillus spp. from annual crops as potential biological control agents of black pod rot of cacao. Biological Control, 46(1), 46–56. https://doi.org/10.1016/j.biocontrol.2008.01.022

73

Michalk, D. L., Badgery, W. B., & Kemp, D. R. (2017). Balancing animal, pasture and environmental outcomes in grazing management experiments. Animal Production Science, 57(9), 1775. https://doi.org/10.1071/AN16132

Miransari, M., Bahrami, H. A., Rejali, F., & Malakouti, M. J. (2009). Effects of arbuscular mycorrhiza, soil sterilization, and soil compaction on wheat (Triticum aestivum L.) nutrients uptake. Soil and Tillage Research, 104(1), 48– 55. https://doi.org/10.1016/j.still.2008.11.006

Moe, K., Mg, K. W., Win, K. K., & Yamakawa, T. (2017). Effects of Combined Application of Inorganic Fertilizer and Organic Manures on Nitrogen Use and Recovery Efficiencies of Hybrid Rice (Palethwe-1). American Journal of Plant Sciences, 08(05), 1043. https://doi.org/10.4236/ajps.2017.85069

Mollet, C., Drancourt, M., & Raoult, D. (1997). RpoB sequence analysis as a novel basis for bacterial identification. Molecular Microbiology, 26(5), 1005–1011. https://doi.org/10.1046/j.1365-2958.1997.6382009.x

Montero, E., Cabot, C., Poschenrieder, C., & Barcelo, J. (1998). Relative importance of osmotic-stress and ion-specific effects on ABA-mediated inhibition of leaf expansion growth in Phaseolus vulgaris. Plant, Cell & Environment, 21(1), 54– 62. https://doi.org/10.1046/j.1365-3040.1998.00249.x

Morella, N. M., Weng, F. C.-H., Joubert, P. M., Metcalf, C. J. E., Lindow, S., & Koskella, B. (2020). Successive passaging of a plant-associated microbiome reveals robust habitat and host genotype-dependent selection. Proceedings of the National Academy of Sciences, 117(2), 1148–1159. https://doi.org/10.1073/pnas.1908600116

Needham, P. (1973). Nutritional disorders. The UK Tomato Manual. Grower Books, London.

Niu, B., Paulson, J. N., Zheng, X., & Kolter, R. (2017). Simplified and representative bacterial community of maize roots. Proceedings of the National Academy of Sciences, 114(12), E2450–E2459. https://doi.org/10.1073/pnas.1616148114

Ogier, J.-C., Pages, S., Galan, M., Barret, M., & Gaudriault, S. (2019). RpoB, a promising marker for analyzing the diversity of bacterial communities by amplicon sequencing. BioRxiv, 626119. https://doi.org/10.1101/626119

Overmann, J., Abt, B., & Sikorski, J. (2017). Present and Future of Culturing Bacteria. Annual Review of Microbiology, 71(1), 711–730. https://doi.org/10.1146/annurev-micro-090816-093449

74

Özdemir, F., & Arslan, S. (2019). Molecular Characterization and Toxin Profiles of Bacillus spp. Isolated from Retail Fish and Ground Beef. Journal of Food Science, 84(3), 548–556. https://doi.org/10.1111/1750-3841.14445

Palleroni, N. J. (1997). Prokaryotic diversity and the importance of culturing. Antonie Van Leeuwenhoek, 72(1), 3–19. https://doi.org/10.1023/a:1000394109961

Palmieri, D., Vitullo, D., De Curtis, F., & Lima, G. (2017). A microbial consortium in the rhizosphere as a new biocontrol approach against fusarium decline of chickpea. Plant and Soil, 412(1–2), 425–439. https://doi.org/10.1007/s11104- 016-3080-1

Panke-Buisse, K., Poole, A. C., Goodrich, J. K., Ley, R. E., & Kao-Kniffin, J. (2015). Selection on soil microbiomes reveals reproducible impacts on plant function. The ISME Journal, 9(4), 980–989. https://doi.org/10.1038/ismej.2014.196

Park, H. Y., Kim, K. K., Jin, L., & Lee, S.-T. (2006). Microbacterium paludicola sp. Nov., a novel xylanolytic bacterium isolated from swamp forest. International Journal of Systematic and Evolutionary Microbiology, 56(Pt 3), 535–539. https://doi.org/10.1099/ijs.0.63945-0

Pavlova, A. S., Leontieva, M. R., Smirnova, T. A., Kolomeitseva, G. L., Netrusov, A. I., & Tsavkelova, E. A. (2017). Colonization strategy of the endophytic plant growth-promoting strains of Pseudomonas fluorescens and Klebsiella oxytoca on the seeds, seedlings and roots of the epiphytic orchid, Dendrobium nobile Lindl. Journal of Applied Microbiology, 123(1), 217–232. https://doi.org/10.1111/jam.13481

Qingxiang Yang, Hao Zhang, Yuhui Guo, & Tiantian Tian. (2016). Influence of Chicken Manure Fertilization on Antibiotic-Resistant Bacteria in Soil and the Endophytic Bacteria of Pakchoi. International Journal of Environmental Research & Public Health, 13(7), 662. https://doi.org/10.3390/ijerph13070662

Quambusch, M., & Winkelmann, T. (2018). Bacterial Endophytes in Plant Tissue Culture: Mode of Action, Detection, and Control. In V. M. Loyola-Vargas & N. Ochoa-Alejo (Eds.), Plant Cell Culture Protocols (pp. 69–88). Springer. https://doi.org/10.1007/978-1-4939-8594-4_4

R Core Team. (2014). R: A language and environment for statistical computing. R Foundation for Statistical Computing. http://www.R-project.org/

Radford, B. J., Yule, D. F., McGarry, D., & Playford, C. (2001). Crop responses to applied soil compaction and to compaction repair treatments. Soil and Tillage Research, 61(3), 157–166. https://doi.org/10.1016/S0167-1987(01)00194-5 75

Ramasamy, D., Mishra, A. K., Lagier, J.-C., Padhmanabhan, R., Rossi, M., Sentausa, E., Raoult, D., & Fournier, P.-E. (2014). A polyphasic strategy incorporating genomic data for the taxonomic description of novel bacterial species. International Journal of Systematic and Evolutionary Microbiology, 64(2), 384– 391. https://doi.org/10.1099/ijs.0.057091-0

Rasche, F., Velvis, H., Zachow, C., Berg, G., Elsas, J. D. V., & Sessitsch, A. (2006). Impact of transgenic potatoes expressing anti‐bacterial agents on bacterial endophytes is comparable with the effects of plant genotype, soil type and pathogen infection. Journal of Applied Ecology, 43(3), 555–566. https://doi.org/10.1111/j.1365-2664.2006.01169.x

Ravindran, B., Mupambwa, H. A., Silwana, S., & Mnkeni, P. N. S. (2017). Assessment of nutrient quality, heavy metals and phytotoxic properties of chicken manure on selected commercial vegetable crops. Heliyon, 3(12), e00493. https://doi.org/10.1016/j.heliyon.2017.e00493

Raymond, B., Johnston, P. R., Nielsen-LeRoux, C., Lereclus, D., & Crickmore, N. (2010). Bacillus thuringiensis: An impotent pathogen? Trends in Microbiology, 18(5), 189–194. https://doi.org/10.1016/j.tim.2010.02.006

Reasoner, D. J., & Geldreich, E. E. (1985). A new medium for the enumeration and subculture of bacteria from potable water. Applied and Environmental Microbiology, 49(1), 1–7.

Ren, G., Zhang, H., Lin, X., Zhu, J., & Jia, Z. (2015). Response of leaf endophytic bacterial community to elevated CO2 at different growth stages of rice plant. Frontiers in Microbiology, 6. https://doi.org/10.3389/fmicb.2015.00855

Riegel, C., & Noe, J. P. (2000). Chicken Litter Soil Amendment Effects on Soilborne Microbes and Meloidogyne incognita on Cotton. Plant Disease, 84(12), 1275– 1281. https://doi.org/10.1094/PDIS.2000.84.12.1275

Romero, F. M., Marina, M., & Pieckenstain, F. L. (2014). The communities of tomato Solanum lycopersicum L.) leaf endophytic bacteria, analyzed by 16S-ribosomal RNA gene pyrosequencing. FEMS Microbiology Letters, 351(2), 187–194. https://doi.org/10.1111/1574-6968.12377

Rosenblueth, M., & Martínez-Romero, E. (2006). Bacterial Endophytes and Their Interactions with Hosts. Molecular Plant-Microbe Interactions®, 19(8), 827– 837. https://doi.org/10.1094/MPMI-19-0827

76

Roy, S., & Kashem, M. A. (2014). Effects of Organic Manures in Changes of Some Soil Properties at Different Incubation Periods. Open Journal of Soil Science, 2014. https://doi.org/10.4236/ojss.2014.43011

Sainju, U. M., Dris, R., & Singh, B. (2003). Mineral nutrition of tomato. 9.

Sanders, E. R., & Miller, J. H. (2010). I, Microbiologist: A Discovery-Based Course In Microbial Ecology and Molecular Evolution. ASM Press.

Santos, S. R., & Ochman, H. (2004). Identification and phylogenetic sorting of bacterial lineages with universally conserved genes and proteins. Environmental Microbiology, 6(7), 754–759. https://doi.org/10.1111/j.1462-2920.2004.00617.x

Schiere, H., & Kater, L. (2001). Mixed crop-livestock farming. A review of traditional technologies based on literature and field experience. FAO Animal Production and Health Paper (FAO). https://agris.fao.org/agris- search/search.do?recordID=XF2003410307

Serrasolsas, I., & Khanna, P. K. (1995). Changes in heated and autoclaved forest soils of S.E. Australia. I. Carbon and nitrogen. Biogeochemistry, 29(1), 3–24. https://doi.org/10.1007/BF00002591

Sexton, W. K., Fidero, M., Spain, J. C., Jiang, L., Bucalo, K., Cruse-Sanders, J. M., & Pullman, G. S. (2020). Characterization of endophytic bacterial communities within greenhouse and field-grown rhizomes of three rare pitcher plant species (Sarracenia oreophila, S. leucophylla, and S. purpurea spp. Venosa) with an emphasis on nitrogen-fixing bacteria. Plant and Soil, 447(1), 257–279. https://doi.org/10.1007/s11104-019-04372-8

Shahzad, R., Khan, A. L., Bilal, S., Asaf, S., & Lee, I.-J. (2017). Plant growth- promoting endophytic bacteria versus pathogenic : An example of Bacillus amyloliquefaciens RWL-1 and Fusarium oxysporum f. sp. lycopersici in tomato. PeerJ, 5, e3107. https://doi.org/10.7717/peerj.3107

Shahzad, R., Khan, A. L., Bilal, S., Waqas, M., Kang, S.-M., & Lee, I.-J. (2017). Inoculation of abscisic acid-producing endophytic bacteria enhances salinity stress tolerance in Oryza sativa. Environmental and Experimental Botany, 136, 68–77. https://doi.org/10.1016/j.envexpbot.2017.01.010

Sharp, R. E., & LeNoble, M. E. (2002). ABA, ethylene and the control of shoot and root growth under water stress. Journal of Experimental Botany, 53(366), 33–37. https://doi.org/10.1093/jexbot/53.366.33

77

Shi, Y., Lou, K., & Li, C. (2010). Growth and photosynthetic efficiency promotion of sugar beet (Beta vulgaris L.) by endophytic bacteria. Photosynthesis Research, 105(1), 5–13. https://doi.org/10.1007/s11120-010-9547-7

Sistani, K. R., Torbert, H. A., Way, T. R., Bolster, C. H., Pote, D. H., & Warren, J. G. (2009). Broiler Litter Application Method and Runoff Timing Effects on Nutrient and Escherichia coli Losses from Tall Fescue Pasture. Journal of Environmental Quality, 38(3), 1216–1223. https://doi.org/10.2134/jeq2008.0185

Szymańska, S., Borruso, L., Brusetti, L., Hulisz, P., Furtado, B., & Hrynkiewicz, K. (2018). Bacterial microbiome of root-associated endophytes of Salicornia europaea in correspondence to different levels of salinity. Environmental Science and Pollution Research, 25(25), 25420–25431. https://doi.org/10.1007/s11356-018-2530-0

Tamura, K., & Nei, M. (1993). Estimation of the number of nucleotide substitutions in the control region of mitochondrial DNA in humans and chimpanzees. Molecular Biology and Evolution, 10(3), 512–526. https://doi.org/10.1093/oxfordjournals.molbev.a040023

Tamura, Koichiro, Nei, M., & Kumar, S. (2004). Prospects for inferring very large phylogenies by using the neighbor-joining method. Proceedings of the National Academy of Sciences of the United States of America, 101(30), 11030–11035. https://doi.org/10.1073/pnas.0404206101

Tariq, M., Hameed, S., Yasmeen, T., Zahid, M., & Zafar, M. (2014). Molecular characterization and identification of plant growth promoting endophytic bacteria isolated from the root nodules of (Pisum sativum L.). World Journal of Microbiology and Biotechnology, 30(2), 719–725. https://doi.org/10.1007/s11274-013-1488-9

Thomas, P., & Upreti, R. (2016). Evaluation of tomato seedling root-associated bacterial endophytes towards organic seedling production. Organic Agriculture, 6(2), 89–98. https://doi.org/10.1007/s13165-015-0111-9

Tian, B.-Y., Cao, Y., & Zhang, K.-Q. (2015). Metagenomic insights into communities, functions of endophytes and their associates with infection by root-knot nematode, Meloidogyne incognita, in tomato roots. Scientific Reports, 5(1). https://doi.org/10.1038/srep17087

Tracy, S. R., Black, C. R., Roberts, J. A., Sturrock, C., Mairhofer, S., Craigon, J., & Mooney, S. J. (2012). Quantifying the impact of soil compaction on root system architecture in tomato (Solanum lycopersicum) by X-ray micro-computed

78

tomography. Annals of Botany, 110(2), 511–519. https://doi.org/10.1093/aob/mcs031

Vos, M., Quince, C., Pijl, A. S., de Hollander, M., & Kowalchuk, G. A. (2012). A comparison of rpoB and 16S rRNA as markers in pyrosequencing studies of bacterial diversity. PloS One, 7(2), e30600. https://doi.org/10.1371/journal.pone.0030600

Wagner, M. R., Lundberg, D. S., Del Rio, T. G., Tringe, S. G., Dangl, J. L., & Mitchell- Olds, T. (2016). Host genotype and age shape the leaf and root microbiomes of a wild perennial plant. Nature Communications, 7, 12151. https://doi.org/10.1038/ncomms12151

Wang, E. T., Tan, Z. Y., Guo, X. W., Rodríguez-Duran, R., Boll, G., & Martínez- Romero, E. (2006). Diverse endophytic bacteria isolated from a leguminous tree Conzattia multiflora grown in Mexico. Archives of Microbiology, 186(4), 251– 259. https://doi.org/10.1007/s00203-006-0141-5

Wang, Q., Garrity, G. M., Tiedje, J. M., & Cole, J. R. (2007). Naïve Bayesian Classifier for Rapid Assignment of rRNA Sequences into the New Bacterial Taxonomy. Applied and Environmental Microbiology, 73(16), 5261–5267. https://doi.org/10.1128/AEM.00062-07

Wang, S.-S., Liu, J.-M., Sun, J., Sun, Y.-F., Liu, J.-N., Jia, N., Fan, B., & Dai, X.-F. (2019). Diversity of culture-independent bacteria and antimicrobial activity of culturable endophytic bacteria isolated from different Dendrobium stems. Scientific Reports, 9. https://doi.org/10.1038/s41598-019-46863-9

Weisburg, W. G., Barns, S. M., Pelletier, D. A., & Lane, D. J. (1991). 16S ribosomal DNA amplification for phylogenetic study. Journal of Bacteriology, 173(2), 697–703. https://doi.org/10.1128/jb.173.2.697-703.1991

Wicaksono, W. A., Jones, E. E., Casonato, S., Monk, J., & Ridgway, H. J. (2018). Biological control of Pseudomonas syringae pv. Actinidiae (Psa), the causal agent of bacterial canker of kiwifruit, using endophytic bacteria recovered from a medicinal plant. Biological Control, 116, 103–112. https://doi.org/10.1016/j.biocontrol.2017.03.003

Wickham, H. (2016). ggplot2: Elegant Graphics for Data Analysis. Springer-Verlag.

Wickham, Hadley, Hester, J., & Francois, R. (2018). readr: Read Rectangular Text Data. (R package version 1.3.1.) [Computer software]. https://CRAN.R- project.org/package=readr

79

Wolf, D. C., Dao, T. H., Scott, H. D., & Lavy, T. L. (1989). Influence of Sterilization Methods on Selected Soil Microbiological, Physical, and Chemical Properties. Journal of Environmental Quality, 18(1), 39–44. https://doi.org/10.2134/jeq1989.00472425001800010007x

Xia, Y., DeBolt, S., Dreyer, J., Scott, D., & Williams, M. A. (2015). Characterization of culturable bacterial endophytes and their capacity to promote plant growth from plants grown using organic or conventional practices. Frontiers in Plant Science, 6. https://doi.org/10.3389/fpls.2015.00490

Xu, W., Wang, F., Zhang, M., Ou, T., Wang, R., Strobel, G., Xiang, Z., Zhou, Z., & Xie, J. (2019). Diversity of cultivable endophytic bacteria in mulberry and their potential for antimicrobial and plant growth-promoting activities. Microbiological Research, 229, 126328. https://doi.org/10.1016/j.micres.2019.126328

Yamamoto, S., & Harayama, S. (1995). PCR Amplification and Direct Sequencing of gyrB Genes with Universal Primers and Their Application to the Detection and Taxonomic Analysis of Pseudomonas putida Strains. APPL. ENVIRON. MICROBIOL., 61, 7.

Yang, Q., Ren, S., Niu, T., Guo, Y., Qi, S., Han, X., Liu, D., & Pan, F. (2014). Distribution of antibiotic-resistant bacteria in chicken manure and manure- fertilized vegetables. Environmental Science and Pollution Research International; Heidelberg, 21(2), 1231–1241. http://dx.doi.org.proxy.lib.ohio- state.edu/10.1007/s11356-013-1994-1

Yanti, Y., Warnita, & Reflin. (2019). Involvement of Jasmonic Acid in the Induced Systemic Resistance of Tomato against Ralstonia syzigiisub sp. Indonesiensis by Indigenous Endophyte Bacteria. IOP Conference Series: Earth and Environmental Science, 347, 012024. https://doi.org/10.1088/1755- 1315/347/1/012024

Yoon, J.-H., Kang, S.-J., Schumann, P., & Oh, T.-K. (2007). Cellulosimicrobium terreum sp. Nov., isolated from soil. International Journal of Systematic and Evolutionary Microbiology, 57(Pt 11), 2493–2497. https://doi.org/10.1099/ijs.0.64889-0

Zhang, B.-H., Salam, N., Cheng, J., Li, H.-Q., Yang, J.-Y., Zha, D.-M., Guo, Q.-G., & Li, W.-J. (2017). Microbacterium lacusdiani sp. Nov., a phosphate-solubilizing novel actinobacterium isolated from mucilaginous sheath of Microcystis. The Journal of Antibiotics, 70(2), 147–151. https://doi.org/10.1038/ja.2016.125

80

Zimmer, W., Hundeshagen, B., & Niederau, E. (1994). Demonstration of the indolepyruvate decarboxylase gene homologue in different auxin-producing species of the Enterobacteriaceae. Canadian Journal of Microbiology, 40(12), 1072–1076. https://doi.org/10.1139/m94-170

81

Tables

Year of Rotation Year of Plots Experiment 2016 2017 2018 2019

2018 Split plot Split plot Split plot R1r5 Sunflowers with pasture with pasture with pasture and P/C and P/C and P/V

Split plot Split plot Split plot R4r3 Sunflowers with pasture with pasture with pasture

and P/C and P/C and P/V

Split plot Split plot Split plot R2r1 Sunflowers with pasture with pasture with pasture

and P/C and P/C and P/V

Split plot Split plot Split plot R3r1 Sunflowers with pasture with pasture with pasture

and P/C and P/C and P/V

2020 Split plot Split plot R3r5 Oats Sunflowers with pasture with pasture and P/C and P/C Split plot Split plot R4r4 with pasture Oats Sunflowers with pasture and P/V and P/C Split plot Split plot R3r2 with pasture Oats Sunflowers with pasture and P/V and P/C Split plot Split plot R4r2 Oats Sunflowers with pasture with pasture and P/C and P/C

Table 2.1. History of management strategy rotations implemented in plots selected for soil sampling of 2018 and 2020 greenhouse experiments. For 2018 and 2019 plots sampled are indicated in bold. P/C: Pasture/chickens. P/V: Pasture/vegetables.

82

Sum of Mean Source of Variation df F Value Pr(>F) Squares Square

Celebrity Aboveground Biomass Treatment 4 680.09 170.02 14.27 5.36x10-5 *** Rep 4 40.89 10.22 0.86 0.51 Residuals 15 178.78 11.92

Height Treatment 4 194.64 48.66 11.93 1.47x10-4 *** Rep 4 21.94 5.49 1.35 0.30 Residuals 15 61.16 4.08

Moneymaker Aboveground Biomass Treatment 4 361.46 90.37 6.36 2.93x10-3 ** Rep 4 7.36 1.84 0.13 0.97 Residuals 16 227.50 14.22

Height Treatment 4 187.74 46.94 10.00 2.97x10-4 *** Rep 4 19.24 4.81 1.03 0.42 Residuals 16 75.06 4.69

*** significant at P=0 ** significant at P=0.001

Table 2.2. Impact of dilution of chicken grazed soils on aboveground biomass and height, as estimated by analysis of variance, of Celebrity and Moneymaker tomato seedlings. Seedlings were grown under greenhouse conditions in chicken-grazed soils diluted with a sterile soil substrate in 2018.

83

Replicates First Inoculation Mean 1 2 3 10-1 TNTC* TNTC* TNTC* 10-2 TNTC* TNTC* TNTC* 10-3 1.7x104 1.3x104 9.5x103 1.3x104 10-4 5.1x103 2.3x103 4.4x103 3.9x103 Sterile Water (Control) 0 0 0 0

Second Inoculation

10-1 TNTC* TNTC* TNTC* 10-2 2.3x104 2.4x104 1.1x104 1.9x104 10-3 3.4x103 3x103 3.8x103 3.4x103 10-4 6x101 3x101 5x101 4.7x101 Sterile Water (Control) 0 0 0 0

*Colony counts above 250 are considered "Too numerous to count" (TNTC)

Table 2.3. Number of CFU/0.1mL for each chicken-grazed soil slurry dilution. The first inoculation of soil slurry was done prior to tomato seed germination. The second inoculation was added to germinated seedlings. .

84

Sum of Mean F Source of Variation df Pr(>F) Squares Square Value Aboveground Biomass Treatment 4 22.83 5.71 1.47 0.24 Block 2 19.23 9.62 2.47 0.10 Rep 7 12.87 1.84 0.47 0.85 Treatment:Block 8 10.74 1.34 0.34 0.94 Residuals 27 105.17 3.90

Height1 Treatment 4 0.07 0.02 4.64 0.01 ** Block 2 2.06x10-3 1.03x10-3 0.27 0.77 Rep 7 0.01 1.67x10-3 0.43 0.87 Treatment:Block 8 0.02 2.28x10-3 0.59 0.78 Residuals 27 0.10 3.84x10-3

Root Biomass Treatment 4 10.25 2.56 1.74 0.17 Block 2 8.01 4.01 2.72 0.08 . Rep 7 4.04 0.58 0.39 0.90 Treatment:Block 8 4.15 0.52 0.35 0.94 Residuals 27 39.79 1.47

Number of Compound Leaves2 Treatment 0.06 . Block 0.42

Root to Shoot Ratio2 Treatment 2.58x10-3 *** Block 0.65

. significant at less than P=0.1 ** significant at P=0.001 1Data was transformed due to non-normality. 2Data was measured using Kruskal-Wallis test.

Table 2.4. Impact of dilution of chicken grazed soils on aboveground biomass, root biomass, height, number of compound leaves and root to shoot ratio of tomato seedlings grown under greenhouse conditions in sterile sand-turface and composted soil substrate with an addition of diluted chicken-grazed soil slurry, for 2020 experiment.

85

Aboveground Number of Root to Shoot Height Root Biomass Biomass Compound Leaves Ratio Treatment Standard Standard Standard Standard Standard Mean Mean Mean Mean Mean Deviation Deviation Deviation Deviation Deviation 10-1 6.31 2.06b 14.45 2.53b 2.82 1.40b 5.00 0.67b 0.42 0.12a 10-2 7.04 1.29ab 14.17 1.27b 3.19 0.85ab 5.78 0.44a 0.45 0.07a 10-3 7.84 2.28ab 15.45 1.69b 3.85 1.09a 5.50 0.71ab 0.50 0.06a 10-4 8.20 1.84a 17.70 2.32a 3.92 1.29a 5.50 0.53ab 0.48 0.09a Sterile Water 6.93 1.46ab (Control) 13.95 1.67b 3.94 0.88a 5.00 0.82b 0.57 0.05b

Block*

B1 6.30 1.55b 14.86 1.97 2.95 1.03b 5.14 0.66 0.46 0.11 B2 7.52 1.80ab 15.30 1.25 3.65 1.16ab 5.40 0.63 0.48 0.10

86 B3 7.76 2.00a 15.28 3.16 3.90 1.17a 5.45 0.76 0.50 0.08

*Block arrangement in the greenhouse: B1 (block closest to the wall of the greenhouse), B2 (block in the middle of B1 and B2) and B3 (block closest to the middle of the greenhouse.)

Table 2.5. Soil dilution treatment and block effect on aboveground biomass, root biomass, height, number of compound leaves and root to shoot ratio of tomato seedlings grown under greenhouse conditions in sterile sand-turface and composted soil substrate with an addition of diluted chicken-grazed soil slurry. Means in a column without a common superscript letter differ (p < 0.05), as analyzed by Fisher’s LSD test.

86

Number of isolates recovered per soil dilutions and tomato varieties 100% 75% 50% 25% 100% Number of Closest match in NCBI Recovered isolate ID1 CGS CGS CGS CGS SSS isolates Identity Database with exact %3 C M C M C M C M C M (Accession number) sequences2 Arthrobacter sp. Anthrobacter 1 (AVC.064) 1 1 (MH699257.1) 98.96% Bacillus 1 (AVC.025) 6 5 3 2 3 4 3 5 1 1 33 Bacillus sp. (MN969919.1) 99.71% Bacillus 2 (AVC.038) 1 1 Bacillus sp. (MH683121.1) 99.71% Bacillus 3 (AVC.051) 1 1 1 1 4 Bacillus sp. (MN865961.1) 99.93% Bacillus 4 (AVC.054) 1 2 5 1 9 Bacillus sp. (MN966966.1) 99.85% Bacillus 5 (AVC.061) 1 1 Bacillus sp. (MF948297.1) 99.28% Bacillus 6 (AVC.063) 1 1 Bacillus sp. (MT355654.1) 99.89% Bacillus 7 (AVC.089) 1 1 Bacillus sp. (MT350130.1) 100% 87

Bacillus 8 (AVC.098) 1 1 Bacillus sp. (MT328683.1) 99.90% Bacillus 9 (AVC.110) 1 1 Bacillus sp. (MN447143.1) 100% Cellulomonas sp. Cellulomonas 1 (AVC.104) 1 1 (GQ369082.1) 100% Cellulosimicrobium sp. Cellulosimicrobium 1 (AVC.023) 1 1 (MN904887.1) 99.84% Chryseobacterium sp. Chryseobacterium 1 (AVC.024) 1 1 (FJ169958.1) 99.41% Enterobacteriaceae 1 (AVC.057) 1 1 Unknown bacteria4 Continued

Table 2.6. Bacterial endophytes recovered from the stem of tomato seedlings grown in chicken-grazed soil dilutions. Identification is based on partial 16S sequence. The number of isolates recovered per each sequence type is shown for each treatment and tomato genotype. Abbreviations: CGS: Chicken-grazed soil, SSS: Sterile soil substrate, C: Celebrity, M: Moneymaker.

87

Table 2.6 Continued

Number of isolates recovered per soil dilutions and tomato varieties 100% 75% 50% 25% 100% Number of Closest match in NCBI Recovered isolate ID1 Identity CGS CGS CGS CGS SSS isolates with Database %3 C M C M C M C M C M exact sequences2 (Accession number) Flavobacterium sp. Flavobacterium 1 (AVC.055) 1 1 2 (MH883902.1) 98.87% Flavobacterium sp. Flavobacterium 2 (AVC.087) 1 1 (MG786610.1) 99.48% Lysinibacillus sp. Lysinibacillus 1 (AVC.045) 1 1 (MN795736.1) 99.42% Lysinibacillus sp. Lysinibacillus 2 (AVC.094) 1 1 (MK414879.1) 100% Microbacterium sp. Microbacterium 1 (AVC.022) 1 1 2 (JQ660113.1) 99.55% 88

Microbacterium sp. Microbacterium 2 (AVC.026) 2 2 (MN923509.1) 99.89% Microbacterium sp. Microbacterium 3 (AVC.027) 1 1 (MK425667.1) 100% Microbacterium sp. Microbacterium 4 (AVC.030) 1 2 3 (MG232347.1) 99.89% Microbacterium sp. Microbacterium 5 (AVC.066) 1 1 (MG016448.1) 100% Microbacterium sp. Microbacterium 6 (AVC.068) 1 1 (KC810832.1) 100% Microbacterium sp. Microbacterium 7 (AVC.093) 1 1 (MH127624.1) 99.78% Continued

88

Table 2.6 Continued

Number of isolates recovered per soil dilutions and tomato varieties 100% 75% 50% 25% 100% Number of Closest match in NCBI Recovered isolate ID1 CGS CGS CGS CGS SSS isolates with Identity Database exact %3 C M C M C M C M C M (Accession number) sequences2 Microbacterium 8 Microbacterium sp. (AVC.105) 1 1 (MG016448.1) 100% Ochrobactrum sp. Ochrobactrum 1 (AVC.033) 1 1 (MG550982.1) 99.91% Paenibacillus sp. Paenibacillus 1 (AVC.056) 1 1 2 (MG963217.1) 99.83% Paenibacillus sp. Paenibacillus 2 (AVC.086) 1 2 3 (AB680470.1) 99.64% 89

Paenibacillus sp. Paenibacillus 3 (AVC.106) 1 1 (MN704412.1) 99.07% Pelomonas sp. Pelomonas 1 (AVC.037) 1 1 (KM253124.1) 99.84% Pseudomonas sp. Pseudomonas 1 (AVC.041) 2 2 (MN326729.1) 99.76% Sphingobium sp. Sphingobium 1 (AVC.065) 1 1 (MF582325.1) 98.10% Sphingobium sp. Sphingobium 2 (AVC.102) 1 1 (MF582325.1) 98.22% Streptomyces sp. Streptomyces 1 (AVC.088) 1 1 (MN704433.1) 100%

89

Closest match in NCBI Length Primer pair Recovered isolate ID1 database (Accession Identity %2 of query3 number)

rpoB-f / rpoB-r Bacillus 3 (AVC.034) Bacillus pumilis (LT906438.1) 100% 562 Bacillus 3 (AVC.039) Bacillus sp. (CP043559.1) 100% 526 Bacillus 3 (AVC.040) Bacillus sp. (CP043559.1) 100% 562 Bacillus 3 (AVC.043) Bacillus sp. (CP043559.1) 100% 575 Bacillus 3 (AVC.048) Bacillus sp. (CP043559.1) 100% 541 Bacillus 4 (AVC.054) Bacillus sp. (CP043559.1) 100% 581 Bacillus 3 (AVC.079) Bacillus sp. (CP043559.1) 100% 575 Bacillus 3 (AVC.099) Bacillus sp. (CP043559.1) 100% 566 Bacillus 3 (AVC.109) Bacillus sp. (CP043559.1) 100% 582 90 Bacillus licheniformis Bacillus 7 (AVC.089) (CP045814.1) 100% 578 Bacillus 8 (AVC.098) Bacillus sp. (CP043559.1) 100% 578

Continued

1AVC: Reference isolates from tomato endophyte culture collection. 2The identity percent describes how similar the query sequence is to the target sequence. The higher the percent identity is, the more significant the match. 3Length of query from recovered isolate.

Table 2.7. Subset of bacterial endophytes recovered from the stem of tomato seedlings grown in chicken-grazed soil dilutions. Identification is based in rpoB gene region.

90

Table 2.7 Continued

Closest match in NCBI database Identity Length of Primer pair Recovered isolate ID1 (Accession number) %2 query3

Univ rpoB-F / Univ rpoB-R

Chryseobacterium carnipullorum Chryseobacterium 1 (AVC.024) (CP033921.1) 91.29% 401 Enterobacteriaceae 1 (AVC.057) Enterobacter cloacae sp. (JX425238.1) 99.74% 389 Flavobacterium 1 (AVC.055) Flavobacterium johnsoniae (CP000685.1) 99.62% 264 Flavobacterium 2 (AVC.087) Flavobacterium johnsoniae (CP000685.1) 99.23% 259 Lysinibacillus 1 (AVC.045) Lysinibacillus sp. (CP026007.1) 99.75% 401

91 Microbacterium 6 (AVC.084) Paenibacillus vulneris (MH087423.1) 98.29% 401

Ochrobactrum 1 (AVC.033) Ochrobactrum anthropi (CP044970.1) 99.75% 395 Paenibacillus 1 (AVC.086) Paenibacillus vulneris (MH087423.1) 98.29% 245 Paenibacillus cellulositrophicus Paenibacillus 2 (AVC.106) (CP045295.1) 94.88% 410 Pelomonas 1 (AVC.037) depolymerans (CP013729.1) 91.77% 414 Pseudomonas 1 (AVC.041) Pseudomonas fluorescens (HM234618.1) 99.24% 394 Pseudomonas 1 (AVC.044) Pseudomonas fluorescens (HM234618.1) 99.24% 396 Sphingobium 1 (AVC.065) Sphingobium herbicidovorans (CP020538.1) 93.39% 243 Sphingobium 2 (AVC.102) Sphingobium herbicidovorans (CP020538.1) 93.70% 254

1AVC: Reference isolates from tomato endophyte culture collection. 2The identity percent describes how similar the query sequence is to the target sequence. The higher the percent identity is, the more significant the match. 3Length of query from recovered isolate.

91

Number of Isolates Soil Slurry Dilution Recovered

10-1 57 10-2 51 10-3 59 10-4 45 Sterile Water (Control) 51

Table 2.8. Number of isolates recovered from surface sterilized stem of tomato seedlings grown across all chicken-grazed soil slurry dilutions.

92

Figures

Figure 2.1. Field layout of the SMSFG plots at Mellinger Research Farm at OARDC, Wooster, Ohio. The 0.6 hectare field is divided in four replicates of five plot treatments installed west of the farmstead.

93

Figure 2.2. Gradient of diluted soil from chicken-grazed pasture plots. Five treatments were generated by diluting soil from chicken-grazed pasture plots with autoclaved soil substrate: 100% chicken-grazed soil, 75%, 50% and 25% dilutions and 100% autoclaved soil substrate.

94

Figure 2.3. Gradient of filtered chicken-grazed soil slurry dilution. Filtered soil slurry was diluted to 10-1, 10-2, 10-3 and 10-4. Sterile water was used as a control.

95

96

Figure 2.4. Differences in aboveground biomass and height of tomato seedlings grown in chicken-grazed soil dilutions after six weeks (n=5, except for Celebrity 75% chicken-grazed soil n=4). A. Aboveground biomass. B. Height. Celebrity and Moneymaker seeds were sowed in a dilution series of chicken-grazed soils using sterile soil substrate. Replicates of each treatment were placed in the greenhouse for six weeks and these were measured before being destructively sampled. ANOVA and Fisher’s LSD tests were calculated, and replicate measurements were graphed as boxplots. Black horizontal lines indicate the median, the lower portion of the box is the first quartile, the upper portion of the box is the third quartile, the whiskers are maximum and minimum, and the outliers are black dots. Uppercase letters indicate significant difference across soil dilutions according to Fisher’s LSD test (p < 0.05) for the Celebrity seedlings and the lowercase letters for Moneymaker seedlings. 2018 experiment.

96

97

Figure 2.5. Total number of bacterial endophytes recovered by genera. Pie chart indicating the percentage recovery of different bacterial genera from tomato seedling stems grown in different soil dilutions (total = 87 isolates).

97

98

Figure 2.6. Impact of dilution of chicken grazed soils on taxa of bacterial endophytes recovered from two tomato varieties, grown in different soil dilution treatments. The bar plot shows the number of bacterial isolates recovered from tomato seedling stems (total = 87 isolates). Different colors indicate the taxa to which each isolate belongs.

98

Figure 2.7. Phylogenetic reconstruction of unique endophytic bacterial taxa recovered from the stem of tomato seedlings based on 16S rRNA gene sequences. Taxa labeled as AVC represent unique taxa recovered in this study and are color coded according to their treatment origin. Reference sequences obtained from the NCBI Database are labeled according to the sample origin. The tree was created using the Maximum Likelihood method based on the Tamura-Nei model and generated in MEGA. The bootstrap consensus tree inferred is from 1000 replicates and branches corresponding to partitions reproduced in less than 60% bootstrap replicates are collapsed. Ferroplasma acidiphilum and Methanobacter cuticularis were used as outgroups. 99

Chapter 3. Characterization of the Endophytic Bacterial Microbiome of Tomato Plants

Grown in a Chicken-Grazed Plots

Introduction

Intensification of cropping systems through increased fertilization and reduced crop diversity negatively affect the abundance of bacteria in soil (Hole et al., 2005;

Postma-Blaauw et al., 2010). On the other hand, practices associated with minimizing the use of off-farm resources, such as organic approaches, reduced to no tillage and crop diversification, are known to promote diversity, and in some cases abundance, of bacteria in soil (Hartman et al., 2018; Hartmann et al., 2015; Peralta et al., 2018). Integrated crop- livestock systems are a type of mixed farming practice where farmers can have different enterprises and save resources by interchanging them on the farm (Schiere & Kater,

2001). These systems provide environmental benefits by implementing diverse cropping systems, incorporating perennial and legume forages and adding animal manure through grazing livestock (Russelle et al., 2007). Most studies on the use of grazing livestock on this type of farming system in the U.S. focus on the benefits of using ruminants (Sulc &

Franzluebbers, 2014). On the other hand, there is a lack of emphasis on the use of chickens for grazing. The use of chickens for meat and egg production as part of crop- 100 livestock systems have been studied mostly outside the U.S. (Bishop, 1995; Devendra &

Thomas, 2002; Udo et al., 2011).

The use of incorporated chicken manure in farming systems, however, has been studied extensively. The chicken manure as a source of nutrients and organic matter, has shown to increase plant yield, suppress disease and improve soil quality and health (Araji et al., 2001; Chellemi et al., 2013; Conn & Lazarovits, 1999; Demir et al., 2010; Riegel &

Noe, 2000). However, concerns have been raised over chicken manure containing human pathogens and antibiotic resistant bacteria (Kyakuwaire et al., 2019). Efforts to mitigate these risks by the use of chicken manure include controlling rate and form of incorporation, composting and mixing with other substrates (Ch’ng et al., 2013; Maru et al., 2015; Ye et al., 2016). Chickens can be integrated to crop-livestock systems by having them graze on pasture plots that can later be used for crops. Pastured poultry is considered a sustainable livestock production system that can be integrated with other farm operations (DEFRA, 2001). This strategy can assist in the incorporation of chicken manure in soil while maintaining chicken meat production.

Tomato (Solanum lycopersicum L.) is one of the world’s most important vegetable crops, surpassing production of 182 million tonnes worldwide in 2018 (FAO,

2018). In 2019, the total value of fresh market tomatoes in the U.S. was estimated at approximately $1.60 billion (USDA-AMS, 2019). Although it is mostly known for its culinary versatility, tomato plants have become a model organism for basic and applied

101 research programs. Scientists use these plants to assess its physiology, breeding, fruit quality and plant-microbe interactions (Bergougnoux, 2014). The bacterial tomato plant microbiome has been explored in efforts to characterize communities and select taxa with functionality important for production (Dong et al., 2019; Ottesen et al., 2013). However, this type of research also unravels the mechanisms for transmission and colonization of bacterial communities from the rhizosphere, to the seed (Bergna et al., 2018; López et al.,

2018; Shaik & Thomas, 2019).

Studies that uncover the richness and diversity of bacterial communities associated to the tomato plant compartments have been more common within the last decade. Most of these studies focus on rhizosphere and root tomato plant microbiome, and how it is affected by soil, plant genotype, and root system architecture (Cheng et al.,

2020; Kwak et al., 2018; Lee et al., 2016; Saleem et al., 2018; Tian et al., 2017). To a lesser extent, there are studies that focus on bacterial communities on tomato seed and aerial parts of the plant, such as leaves. (Morella et al., 2019; Romero et al., 2014; Simon et al., 2001; Toju et al., 2019). Allard et al. (2016), Dong et al. (2019) and Ottesen et al.

(2013) have characterized the rhizosphere, phyllosphere and endosphere of roots, stems, leaves, fruits and seeds of tomato plants grown in greenhouse and field conditions. Still, there is a lack of knowledge on the impact of diverse experimental sites and land- management practices on bacterial community composition and variation between tomato plant tissues. Moreover, understanding the impact of agricultural management strategies can assist in modulating the plant microbiome to promote plant health.

102

Studies of integrated crop–livestock systems using chicken grazing impacting the bacterial plant microbiome remain sparse. The aim of this chapter is to characterize the endophytic bacteria of tomato plants grown in pasture plots a year after chicken grazing.

Culture-dependent and culture-independent techniques will provide insight into the impact of crop-livestock rotation in tomato and soil microbiomes. The objectives are to:

(a) determine the impact of chicken grazed pasture plots (CGPPs) on soil heath and tomato plant tissue biomass, (b) evaluate cultured endophytic bacteria composition obtained from the stem of tomato plants grown in CGPPs and (c) using amplicon sequencing, determine the effect of chicken grazing history on bacterial communities of different tomato plant tissues and soil. We hypothesize that bacterial endophyte diversity is higher in plants grown in CGPPs, as opposed to those grown in pasture only plots. To test this hypothesis, a field trial was conducted in the Summer of 2019 where tomato plants were grown in soils with differing chicken grazing history (Figure 3.1). During the field trial, tomato plant tissues (stem, leaves, roots, fruit) and rhizosphere soil were collected at two time points: before flowering and at harvest (Figure 3.2). This project will help acquire a greater understanding of the impact of agricultural management strategies that increase soil quality on the microbiome of this highly demanded commodity.

103

Materials and Methods

Field Site Description

In the year 2016, the project Diversification of Small and Medium Sized Farms

Grant (SMSFG) was established at Mellinger Research Farm at the Ohio Agricultural

Research and Development Center (OARDC), The Ohio State University (OSU). The

0.61-hectare experimental field includes integrated crop and livestock systems as part of a rotation and its impact is evaluated on subsequent vegetable yield (Figure 2.1). Each plot undergoes a management strategy with rotations every year. For this dissertation, we focused on the pasture and chicken grazing rotation (split plot pasture and pasture chicken). The plots were 105 (R1r5), 403 (R4r3), 201 (R2r1) and 301 (R3r1). In 2017, all plots were planted with a mixture of grass and clover for first year establishment and were mowed prior to the growing season for the second-year pasture in 2018. The composition of the seed mixture for pasture plots was 13.61 kg of HQ-R Pasture Mix

(JDS Seeds Orrville, OH) with 2.27 kg of white (JDS Seeds Orrville, OH) and red clover

(RKO). HQ-R Pasture Mix contains 29% Niva orchard grass, 29% Laura meadow fescue,

20% premium perennial rye grass, 10% Calibra perennial rye grass and 10% Climax timothy.

Chicken grazing was setup as split-plot within each pasture plot. For both years

(2017 and 2018) broiler chickens were placed in 1.5 m x 3.5 m tractors and allowed to

104 graze in pasture plots. In August 2017 and 2018, tractors were moved once daily across the selected pasture split plots for 37 and 32 days, respectively. Chickens grazed the mixture of available pasture and chicken diet was supplemented with feed. Pasture/ pasture-chicken split plots were terminated on May 6th 2019 with application of

Roundup® with Ammonium sulfate. On May 16th 2019 plots were tilled prior to preparing vegetable beds for planting. Between May 20th and May 30th, landscaping fabric and drip-tape were layed on all four plots, and plots were fenced. A mixture of pasture and clover seeds was hand spread in the exposed areas between vegetable beds.

During the same week, other vegetables were planted in addition to the tomatoes for this experiment. For each split plot, two 2.3 m rows were used to plant tomato seedlings.

Preparation of Tomato Seedlings for Transplant

On May 20th, 6 L of soil was collected from each CGPP half and pasture only

(POP) half from split plots 105, 403, 201 and 301. The shovel was cleaned with ethanol between each half and each plot. Soil was bagged in separate plastic bags per split plot. In the greenhouse, soil was poured into plastic containers, separated by treatment and plot, and mixed-together and homogenized by hand. This soil was used as substrate for germination of tomato plants for the field experiment to avoid introduction of endophytes from potting mix (Figure 3.3).

105

The field soil was added to twelve 10.16 cm pots, representing each split plot

(CGPP and POP) in the field (four field replications), for a total of 96 pots. The remaining soil was stored in plastic bags and paper bags. Plastic bags were placed in a -

80oC freezer and paper bags were left open inside the greenhouse for drying. One

Celebrity (Johnny’s Selected Seeds) tomato seed was added for each pot. Pots were evaluated for germination 12 days after sowing and those that did not have germination, had a new seed planted. No significant differences in seed germination were observed.

Tomato seedlings were grown in the greenhouse for three weeks. Greenhouse conditions ranged from a maximum of 35oC and a minimum of 18oC. On June 14th, 10 tomato seedlings were selected per treatment and plot to be transplanted to the corresponding split plot were soil was originally collected from (Figure 3.4). For each split plot, two 2.3 m rows were used to plant tomato seedlings. Each row had five tomato seedlings planted

70 cm apart. After each seedling was planted, water and Bonide® Slug Killer were added around the bottom of the stem. On July 17th, tomato plants were staked by using bamboo stakes and twine and further checked throughout the rest of the season for growth and disease.

Soil Health Measurements

Air dried soil samples from each split plot were sieved to 2mm. Soil samples were analyzed for permanganate oxidizable carbon (POXC), mineralizable carbon and soil protein by the Soil Fertility Lab in the School of Environment and Natural Resources at

106 the OARDC (Soil Fertility Lab, 2019). POXC is a measurement for readily available carbon. Soil protein is a measurement of readily-available nitrogen (Hurisso et al., 2018).

Respiration is a measurement of mineralizable carbon by microbial activity by rewetting dried soil and incubating for 24 hours (Franzluebbers et al., 2000). These measurements were obtained to compare the effect of chicken grazing prior to planting of the tomato on the active pool of organic matter in soil which is rapidly cycled and acquired by crops

(Wander et al., 1994; Woods, 1989). Therefore, they are sensitive indicators of fertility for recently altered management practices (McCarty et al., 1998). Larger active carbon pools reflect greater soil organic matter and enhanced soil fertility (Wander, 2004).

Tomato Plant Sampling and Evaluation

Tomatoes were sampled at two time points, for both biomass and endophyte characterization: July 5th and September 5th.

On July 5th, when the plants were about to the reach the flowering stage (around five compound leaves per plant), four tomato plants were sampled for each split plot. A shovel was introduced in the soil 15.24 cm away from the stem and around the plant to loosen the soil around the roots. The whole plant was removed from the plot and placed in a paper bag. Aboveground plant tissues were separated from the below-ground by cutting with scissors at the base of the stem. Aboveground plant tissues were measured for biomass and number of compound leaves.

107

On September 5th, when the plant had mature fruits, four tomato plants were sampled for each split plot, as described above. However, in September, all plants were infected with septoria leaf spot of tomato, caused by Septoria lycopersici. Plants were diagnosed by the Vegetable Pathology Lab at OARDC. Some plants had leaves severely dried up and the compound leaf count considered the leaf node. Aboveground plant tissues were separated from the below-ground by cutting with scissors at the base of the stem. Aboveground plant tissues, including tomatoes, were measured for biomass, number of compound leaves, number of symptomatic tomatoes and number of total tomatoes harvested.

Plant Tissue Sampling, Surface Sterilization and Isolation of Endophytic Bacteria from Tomato Plant Stem

For both July and September tomato plants, individual plant tissues were sampled.

First, stems were separated from leaves and excess soil loosely attached was removed from roots. Rhizosphere soil (soil attached to the roots) was collected by shaking roots gently over aluminum foil. Root, stems and leaves were surface sterilized by immersing tissue in 70% ethanol (1 min), 20% bleach mixed with 0.1% Tween (1 min), 70% ethanol

(1 min) and rinsed three times with sterile distilled water. One healthy fruit from one plant per split plot was surface sterilized by immersing in 70% ethanol and drying in a laminar flow hood for 30 minutes. After, fruit flesh was cut open using a sterile scalpel

108 and seeds were removed. Seeds were placed on paper, dried in laminar flow hood and stored in paper envelopes.

For culturing of bacterial endophytes, one healthy (or healthier) plant per split plot was selected. The edges of the surface sterilized stem were cut off with a sterile scalpel, and the stem and tomato, was cut into several 1-cm section pieces. Tissue sections were plated on 1/10 TSA and yeast dextrose carbonate agar and incubated at ambient temperature (approximately 22oC) for 48 hours. Emerging bacterial colonies from the stem and tomato flesh sections were transferred up to five times to obtain pure cultures. Remaining tomato flesh, stem, leaves, roots and rhizosphere soil were stored in aluminum foil and placed in -80oC freezer.

Cultured Endophyte DNA Extraction, PCR Amplification and Sequencing of the 16S rRNA Gene Region

The 16S rRNA Gene was used to characterize bacterial endophytes recovered from tomato tissues. A boiling lysis method, according to Sanders and Miller (2010), was used to extract genomic DNA from individual bacterial isolates. A single colony was selected and immersed in 200µL of sterile water in a 1.5mL tube. The tube was then incubated in a water bath at 100oC for 10 minutes, transferred to ice for 10 minutes followed by centrifugation at 10,000 rpm for 10 minutes. 150µL of the supernatant containing the genomic DNA was transferred to a new 1.5mL tube. PCR amplification of the 16SrRNA gene for bacterial endophytes was performed using primers 8F (5’- 109

TGGAGAGTTTGATCCTGGCTCAG-3’) and 1492R (5'-

GGTTACCTTGTTACGACTT-3') (Weisburg et al., 1991). The PCR reaction mixture consisted of 2 μL of DNA template, 1× GoTaq Flexi Buffer (Promega), 0.2 mM dNTP mix (Promega), 2 mM of MgCl2 (Promega),1 mg/ml of Bovine serum albumin (BSA),

0.03 GoTaq DNA polymerase (Promega), 0.5 μM each primer, and molecular grade water (Hyclone). PCR reaction conditions were an initial denaturation at 95oC for 5 minutes, 35 cycles of denaturation at 94oC for 1 minute, annealing at 54oC for 45 seconds and elongation at 72oC for 1 minute and a final elongation cycle of 72oC for 8 minutes.

5µL of each amplification product was analyzed by agarose gel (1%, containing 5 μL of

10,000X SYBR Safe, Thermo Fisher Scientific) electrophoresis in 0.5 Tris borate EDTA

(TBE) buffer with 100 bp Plus DNA ladder (Promega). PCR products were evaluated through gel electrophoresis. The gels were visualized under UV transillumination (Gel

Logic 112 Imaging System Kodak). Enzymatic PCR cleanup was done by adding 5µL of

PCR product, 1µL of Alkaline Phosphatase and 1µL Exonuclease I (Ilustra) into a new

0.2ml tube. PCR cleanup conditions consisted incubating at 37°C for 15 minutes followed by incubating at 80°C for 15 minutes. Preparation of cleaned PCR products for sequencing consisted of adding 10.75µL of clean PCR product and 1.25µL of 6.4 pmol of one of two primers previously used for PCR reaction. Samples were sent to sequence at the Genomics Shared Resource at The Ohio State University Comprehensive Cancer

Center, Columbus, OH. Partial nucleotide sequence was determined with Sanger sequencing using forward and reverse primers.

110

Isolate Sequence Data Analysis

For individual sequences, chromatograms were visualized and trimmed for quality using 4Peaks software (Gerberastraat & Aalsmeer, 2001). Forward and reverse reads for each isolate were merged with CLC Sequence Viewer (Genomics Workbench Version

20.0, 2020). Sequences were compared to the NCBI (National Center of Biotechnology

Information) nucleotide database using the basic local alignment search tool (BLAST, blastn) (Altschul et al., 1990), and the Ribosomal Database Project (RDP) (Wang et al.,

2007) Classifier tool. Recovered isolate sequences were aligned and pairwise distance was computed using the Maximum Composite Likelihood model (0.000; sequences had

100% nucleotide identity) to identify unique isolates (Tamura et al., 2004).

Culture-Independent Characterization of Bacterial Endophytes

Total DNA was extracted from the different sampled sources. DNA extraction of all plant tissue (stem, leaf, roots, tomato flesh and tomato seeds), chicken manure and rhizosphere soil samples were processed with the DNAeasy Plant Mini Kit (Qiagen) according to the manufacturer’s instructions, but with the following tissue specific modifications (for grinding and maceration). For tomato tissue, surface sterilized stem, leaf, root and tomato samples were macerated using mortar and pestle. Tomato seed samples were submerged in liquid nitrogen and macerated using mortar and pestle. Prior to extraction, bulk and rhizosphere soil samples were dried using a centrifugal evaporator

111 for at 60°C 30 minutes. Chicken manure samples were lyophilized and macerated by adding sample into a 2 ml tube with two stainless steel beads and then the tube was vortexed for 30 seconds at 4 M/S high speed by a high-speed benchtop homogenizer

(FASTPREP-24™ 5G, MP BIO). DNA extraction of bulk soil samples was processed with SurePrep Soil DNA Isolation Kit (Fisher BioReagents). Negative extraction controls were processed for each kit.

To assess the DNA quantity and purity, 3µL of each DNA sample was analyzed by agarose gels (1%, containing 5 μL of 10,000X SYBR Safe, Thermo Fisher Scientific) electrophoresis in 0.5 TBE buffer at 70 mV for 90 min with GeneRuler 1 kb Plus DNA

Ladder (Thermo Fisher Scientific). The gels were visualized under UV transillumination

(Gel Logic 112 Imaging System Kodak). ImageJ (Schneider et al., 2012) was used to determine relative PCR product concentration from gel images, using Gene Ruler ladder as a reference. DNA concentration was adjusted to 5 ng/μL.

Samples were submitted to the Molecular, Cellular and Imaging Center (MCIC) at

OARDC, Wooster, OH for preparation and sequencing of 16S rRNA Amplicon libraries.

There, DNA fragments of the bacterial 16S rRNA gene were amplified using a modified semi-nested PCR previously described to limit amplification of plant products (Beckers et al., 2016). The first PCR was run with the primers 799F (5′-

AACMGGATTAGATACCCKG-3′) and 1391R (5’-GACGGGCGGTGWGTRCA-3′), for 15 cycles (Chelius & Triplett, 2001; Walker & Pace, 2007). Then, a 1:20 dilution of

112

PCR product was used as template for the second PCR with pair 967F (5′-

AACMGGATTAGATACCCKG -3′) and 1391R (5′-GACGGGCGGTGWGTRCA-3′) fused with Illumina MiSeq adapters (Sogin et al., 2006). The second PCR was run for 10 of cycles. Individual sample indexes were added through a dual-indexing approach with a final 8 cycles. PCR products were then pooled in equimolar concentration and the pooled library was size selected using Blue Pippin Prep (Sage Science) with a 2% agarose gel cassette and set to collect products of ~ 600 bp, to enrich for bacteria (and exclude plant mitochondrial amplification with a band of ~ 1000 bp with Illumina adapter). Amplicon libraries were then sequenced in a MiSeq 2x300 (Illumina) platform. A total of 154 samples were sequenced in one sequencing run, including 149 field samples and 5 DNA extraction controls.

Amplicon Sequence Data Processing and Analysis

Amplicon sequence data was processed and analyzed (Figure 3.5). Primer and adapter trimming for raw paired reads were done using Cutadapt version 1.18 (Martin,

2011). R (R Core Team, 2014) was used to perform the sequence data processing and statistical analysis. The “DADA2” package (Callahan et al., 2016) was used to trim and filter reads, infer sequence variants, and merge the resulting sequence variants to obtain amplicon sequence variants (ASVs). ASVs were aligned to the SILVA reference database version 132 release (Quast et al., 2013) and taxonomy was assigned. Following sequence alignment using the “DECIPHER” package (Wright, 2016), phylogenetic relatedness of

113 all ASVs was estimated by fitting a generalized time-reversable, with Gamma rate variation (GTR+G+I), maximum likelihood tree using the "phangorn" package (Schliep et al., 2017).

Amplicon Data Analysis and Statistics

Data was imported to “phyloseq” (McMurdie & Holmes, 2013) and was analyzed as a “phyloseq” object to explore ecological metrics and identify differentially abundant taxa. The “decontam” package was used to determine any possible contaminants in the samples submitted for sequencing (Davis et al., 2018). Using this package, the abundance of ASVs considered contaminants were then compared to those in control samples. None of the ASVs tagged as contaminants overlapped with ASVs in control samples, which suggest they were real reads. Sequences matching “Chloroplast”, “Mitochondria” and

“Eukaryota” were removed. Using the “ampvis2” package (Andersen et al., 2018), rarefaction curves were generated to estimate sampling efforts. Samples from different origins were analyzed separately due to differences in sequencing depth, differences in sample handling, DNA extraction procedures, and expected differences in DNA extraction efficiency for each sample type. Cumulative Sum Scaling (CSS) was used as a normalization method for our data due to its ability in detecting differences in the rare members of the communities (McKnight et al., 2019). CSS normalization was performed using the “metagenomeSeq” package (Paulson et al., 2013). To observe distribution of relative abundance of ASVs per sample at phylum level, plot abundances were generated

114 using the “phyloseq” package. Three different alpha diversity metrics were selected to interpret bacterial community compositions: Chao1 Species Richness, Shannon's

Diversity Index and Faith's Phylogenetic Diversity. Chao1 index is used to estimate bacterial richness (abundance of ASVs in a given sample) (Chao et al., 2006). Shannon's

Index is used to estimate bacterial richness as well as evenness of the community

(frequency of ASVs in a given sample) (Shannon, 1948). Faith's Phylogenetic Diversity addresses biodiversity using phylogenetic information (Faith, 2018). Chao1 Species

Richness and Shannon's Diversity Index were estimated using the “breakaway” package

(Willis & Bunge, 2015). Faith's Phylogenetic Diversity was estimated using the

“metagMisc” package (McMurdie & Holmes, 2013; Oksanen et al., 2019). Using the

“agricolae” package (de Mendiburu, 2020), Kruskal-Wallis test was used to determine significant index differences.

Due to variability of diversity indexes and sample size, stem, root and rhizosphere samples were the only ones selected for beta diversity analysis. For beta diversity analysis, Principal Coordinates Analysis (PCoA) was performed based on weighted

UniFrac distance matrices, to observe changes in taxon abundance and taxa distribution between treatments, sample origin and sampling month (Lozupone & Knight, 2005). A two-dimensional ordination plane was plotted with the “phyloseq” package (McMurdie &

Holmes, 2013) and the “ampvis2” package (Andersen et al., 2018). The statistical significance of the differences in bacterial community composition between sampling month (July and September), sample origin (root, stem and rhizosphere soil), and

115 treatment (chicken grazing and pasture only) was assessed through Permutational

Multivariate Analysis of Variance (PERMANOVA) with 1000 permutations (Anderson,

2017). PERMANOVA was performed using the “vegan” package (Oksanen et al., 2019).

Using “DESeq2” package (Love et al., 2014), differentially abundant taxa were analyzed to observe treatment effects between CGPPs and POPs.

Plant Growth and Soil Health Data Analysis and Statistics

The Shapiro-Wilks and Levene’s tests for normality and homogeneity of variance were run in R (Oksanen et al., 2019) to determine any non-normality and equal variance of sample groups. Log 10 transformation was used on data that did not meet parametric assumptions. Mixed design analysis of variance (ANOVA) was used to determine significant differences in measurements between the treatments and blocks. Kruskal-

Wallis test was performed on data that did not meet parametric assumptions after transformation. Significant F-values (p<0.1) obtained led to using Fisher’s least square difference (LSD) test to determine differences between treatments. R (R Core Team,

2014) was used to perform the statistical analyses using “agricolae” (de Mendiburu,

2020), “car” (Fox & Weisberg, 2019), "lsmeans" (Lenth, 2016) and "readr" (Wickham et al., 2018) packages.

116

Results

Impact of Chicken Grazing History on Soil Health

In May 2019, prior to tomato planting, soil samples were collected to test for permanganate oxidizable carbon (POXC), soil protein and respiration. Soil protein and respiration were shown to be significantly different between treatments at p<0.1 (Table

3.1). The means for POXC, soil protein and respiration measurements of soil from

CGPPs increased by 7.6%, 13.1% and 32%, respectively, compared to POPs.

Effect of Chicken Grazing History on Plant Tissue Measurements of Tomato Plants Sampled in July and September

During the field trial at Mellinger Farm, tomato plant tissues (stem, leaves, roots, fruit) were collected and measured at two time points: before flowering (July 5th) and at harvest (September 5th). For tomato plants collected in July, above-ground biomass and number of compound leaves were measured. In July, no significant differences were observed in tomato biomass between CGP and POP, however significant block effects were observed for aboveground biomass and number of compound leaves (Table 3.2).

Fisher's LSD tests showed that the biomass and number of leaves of plants grown on plots 201 and 301 were significantly higher from those grown in 105 and 403 (Table 3.3).

117

For tomato plants collected in September, no significant differences were observed for any of the tomato measurements, however, the means for plants grown in

CGPPs were larger than for those grown in POPs. In September, there were significant differences in above-ground biomass, total biomass of tomatoes harvested, number of symptomatic tomatoes harvested, number of total tomatoes harvested and root biomass of the blocks where the tomato plants grew (Table 3.4). Fisher's LSD tests showed that the above-ground biomass, number of compound leaves, number of total tomatoes harvested and root biomass plots 201 and 301 were significantly higher from those grown in 105 and 403 (Table 3.5).

Impact of Chicken Grazing History on Cultured Endophytic Bacteria Composition

A total of 124 endophytic bacteria were recovered from surface sterilized stems and tomato flesh of tomato plants from July and September. By computing the pairwise distance of each 124 bacterial sequences, 29 of these represent unique taxa. The classification of these 29 taxa, based on >97% similarity to GenBank database entries queried through blastn is shown in (Table 3.6). Recovered isolates belong to 19 genera within four phyla: Actinobacteria, Firmicutes, Proteobacteria and Bacteroidetes. Of the

124 isolates recovered, the stem community had 90 isolates belonging to 27 taxa; the tomato flesh had 34 isolates belonging to 9 taxa. 14 taxa were recovered exclusively in

September. In July, more than 30% of the bacterial endophytes recovered were Bacillus spp. Pseudomonas spp. compromised 26% of the bacterial endophytes recovered in July. 118

In September, nearly half of the bacterial endophytes recovered were Bacillus spp.

Pseudomonas spp. compromised 27% of the bacterial endophytes recovered in

September. The majority of the taxa found across all treatments, tomato plant tissues and sampling months belong to Bacillus 1, followed by Pseudomonas 1. In July and

September, Bacillus spp. and Pseudomonas spp. were the only genera found on all treatments in all blocks (Figure 3.6 and 3.7). Additionally, taxa belonging to Bacillus represented 67% of the isolates recovered from tomato flesh.

The only genera recovered in both time points and treatments, were Bacillus,

Pseudomonas, Microbacterium, and Pantoea. The rest of the isolates, that form around

25%, were present in one of two time points or treatments. Taxa belonging to Bacillus,

Pseudomonas and Microbacterium were the only ones found on both stem and tomato flesh. Taxa belonging to Comamonas, Massilia and Stenotrophomonas were only recovered from CGPPs. Taxa belonging to the Enterobacteriaceae family were found in both treatments.

Tomato-Associated Bacterial Communities and Diversity Analyses

Due to significant block effect and high incidence of septoria leaf spot on tomato, only plant tissue samples obtained from plots 201 and 301 were used for amplicon sequencing. The DNA sequencing of 16S rRNA gene region from bacterial communities in soil, tomato plant tissues and chicken manure resulted in a total of 8,658,317 reads

119

(Table 3.7). After primer removal, filtering, and removal of chloroplast, mitochondria and eukaryotic reads, 981,385 reads remained with an average of 6,372 reads per sample

(Tables 3.8 and 3.9). 148 samples remained after filtering. Control samples had an average of 16 reads and ASVs found in these samples were not abundant in any of the soil, plant tissue and chicken manure samples. Due to these findings, there were no concerns for contamination of the rest of the samples. Control samples were removed from the dataset for analysis. After removing controls, 18,958 ASVs were identified in

143 samples.

ASV richness was plotted using rarefaction curves by displaying the total reads sequenced for each sample by their origin (Figure 3.8). Rarefaction curves for bulk soil, stem, tomato seed and tomato flesh samples leveled off at less than 8,000 reads.

Rarefaction curves for leaf samples leveled off at less than 2,000 reads. Rhizosphere soil and root samples had a sequencing depth surpassing 30,000 reads and the rarefaction curve leveled off at around 30,000 and 45,000 reads, respectively. Rarefaction curves for chicken manure samples surpassed 20,000 reads, and the curve leveled off below 20,000.

Overall, the curves of samples from rhizosphere soil, root and chicken manure leveled off at a higher sequencing depth, therefore had higher species richness, than the curves of samples from bulk soil, stem, leaf, tomato seed and tomato flesh.

The most abundant phyla in all samples were Proteobacteria, Actinobacteria,

Firmicutes and Bacteroidetes (Figures 3.9-3.13). Proteobacteria dominated more than half

120 of the bacterial normalized reads from the bulk soil, root, stem, leaf, tomato flesh and seed samples (60%, 77%, 77%, 86%, 75% and 58% respectively) (Figures 3.9-3.14).

Actinobacteria was the second most abundant phyla in bulk soil, root, stem, tomato flesh and seed (21%, 10%, 9%, 10% and 22% respectively) (Figures 3.9-3.11, 3.13 and 3.14).

However, rhizosphere soil samples contained an even amount of Proteobacteria (34%) and Actinobacteria (34%) (Figure 3.15). In chicken manure, Firmicutes dominated 82%, followed by Proteobacteria (17%) (Figure 3.16). In addition to phyla, abundance of families were evaluated within each sample origin. Burkholderiaceae was abundant in bulk soil (20%), leaf (11%), rhizosphere soil (4%), root (21%), stem (14%) and tomato flesh (3%). was among most abundant families in bulk soil (3%), leaf (45%), root (6%), seed (5%), stem (25%) and tomato flesh (11%).

Enterobacteriaceae was found in high abundance in all plant tissues, dominating tomato flesh (42%) and tomato seed (37%). Chicken manure was mainly composed of

Lactobacillaceae (56%), Enterobacteriaceae (15%) and Enterococcaceae (10%).

Alpha diversity metrics (Chao1 species richness, Shannon’s diversity index and

Faith’s Phylogenetic diversity) were estimated for bacterial communities from bulk soil, chicken manure, rhizosphere soil and plant tissue samples collected in May, June and

September. (Tables 3.10-3.12). In May and July, there was a trend of higher richness, evenness and phylogenetic diversity for soil and plant tissue samples obtained from

CGPPs. In September, richness, evenness and phylogenetic diversity was lower for all samples obtained from CGPPs, with the exception of seed and leaf samples. The bacterial

121 communities in the rhizosphere soil showed the highest estimates among all of the tested samples across sampling months and treatments. There were significant differences in the diversity of bacteria recovered from rhizosphere soil and leaf samples collected from tomato plants grown between plots 201 and 301 (Table 3.13). Rhizosphere soil samples collected from tomato plants grown in plot 201 had higher species richness than plot 301 in both sampling months and treatments. Leaf samples belonging to CGPP 301 had the highest richness.

Due to variability of diversity indexes and sample size, stem, root and rhizosphere samples were the only ones selected for further analysis. PCoA ordination, from weighted normalized UniFrac distance matrix, was performed to evaluate the effect of sample origin, sampling month and treatment, on bacterial community structure (Figure 3.17).

The first and second principal coordinates explained 47.2% and 20.9% of the total variation, respectively. Bacterial community structure had significant differences in bacterial communities based on sample origin (PERMANOVA, p value = 9.99x10-5) and sampling month (PERMANOVA, p value = 1.40x10-3). The strongest clustering was observed for rhizosphere soil samples followed by stem and root, indicating similarity in bacterial community profiles, regardless of the treatment. Additionally, bacterial communities differ between the stem and root. Another PCoA ordination plot was generated to show the distribution of the five most dominant phyla (Figure 3.18). In this plot, data points representing one of the top five most dominant phyla are plotted within the same ordination space, therefore taxa distribution can be correlated with samples

122 using the same ordination space (Pavoine et al., 2004). The top five phyla (Acidobacteria,

Actinobacteria, Bacteroidetes, Firmicutes and Proteobacteria) were distributed among all samples with the exception of Acidobacteria being the most clustered with rhizosphere samples.

Differentially abundant taxa in tomato tissue, in response to chicken grazing history, were identified by comparing samples of CGPPs and POPs within sampling months (Table 3.14), using DeSEq. A total of 67 ASVs were differentially abundant between treatments, across all tissue samples (Table 3.14, Table A.1). Only one ASV from Actinobacteria found in rhizosphere soil from tomato plants grown in CGPPs had significantly lower relative abundance, compared to POP. The majority of the ASVs shown to be more abundant in CGPPs belonged to Proteobacteria, followed by

Actinobacteria.

Discussion

This study found that plots with history of chicken grazing resulted in significant increases in soil fertility. Although not significant, there was an increasing trend of biomass measurements from tomato plants grown in CGPPs compared to those from

POPs. Taxonomic composition of dominant groups of recovered endophytic bacterial isolates were consistent with those found by amplicon sequencing. Higher amounts of C in CGPPs could favor copiotrophic bacteria, such as dominant phyla in this study, alpha

123

Proteobacteria and Bacteroidetes. A total of 65 ASVs were differentially abundant in soil and tomato plant samples from CGPPs. High abundance of Enterobacteriaceae in the endophytic community of all plant tissues could raise concerns for the use of chicken manure, which is known to harbor human pathogens that can potentially contaminate fresh produce. Isolates recovered from tomato plants grown in CGPPs could possess potentially antagonistic properties towards plant pathogens. Sampling month and origin of sample were the main drivers in bacterial community composition and diversity, regardless of chicken grazing history. These findings indicate that plant- and soil- associated microbiomes might have a stronger response to management strategies in earlier plant development stages.

Soils with a history of chicken grazing had significantly higher microbial activity and readily available nitrogen, although no significant increase in tomato plant biomass.

However, the means for POXC measurements in soil and tomato plant biomass sampled from CGPPs in September are higher than those grown in POPs. The active pools of organic matter measured serve as indicators of available carbon and nitrogen that is eventually taken up by microbes (Sahoo et al., 2019). Measurements for active carbon, mineralizable carbon and soil protein tend to be most sensitive to changes in the environment (Culman et al., 2012; Hurisso et al., 2018; Lai et al., 2012). The practice of chicken grazing might be increasing fertility in soil. Hilimire et al. (2013) conducted a similar study where they compared soil fertility of plots with recent history of pastured poultry, as well as biomass changes of crop plants grown in greenhouse experiment with

124 those soils. Although labile C and N were not measured, pastured poultry plots had significantly higher total C, total N, and organic matter compared to soils without pastured animal inputs. Additionally, biomass and height of sunflowers and beans grown in those soils was greater but varied by farm. Unlike our study, the authors were also able to evaluate the amount of chicken manure incorporated in the soil. Studies evaluating varying application rates of chicken manure have reported variable results (Conn &

Lazarovits, 1999; Maru et al., 2015; Riegel & Noe, 2000). Therefore, while chicken grazing might be beneficial for soil fertility and plant biomass, it is important to know the application rate of chicken manure to further distinguish its effects.

Variations in plant biomass observed might be due to field topography and precipitation during the growing season, regardless of treatment. Tomato plants grown in plots 201 and 301 from both sampling times had a significant increase in biomass compared to other plots. Weather during the 2019 growing season had a high precipitation with a total of 53.44 cm, which is a 51% increase of the average precipitation from other years (CFAES, 2020). Plots 105 and 403 are closest to the north of the field. According to Web Soil Survey, this area has less drainage (Soil Survey Staff,

2019). Therefore, there could be differences in soil compaction and infiltration within the field that can increase flooding in plots where tomato plants that had significantly lower biomass (Alaoui et al., 2018).

125

Dominant phyla (Actinobacteria, Proteobacteria, Bacteroidetes and Firmicutes) of bacterial taxa identified were uniform between culture-dependent and -independent approaches. According to amplicon sequencing data, Proteobacteria dominated more than half of the bacterial community in samples from plant tissue origin. Moreover, cultured endophytes had a predominance of Proteobacteria, as well as Firmicutes. Both phyla have been consistently shown as abundant across tomato plant tissues in other studies (Santoyo et al., 2016; Tian et al., 2017; Xia et al., 2015). Across all treatments, blocks, plant tissue and sampling points, taxa belonging to Bacillus and Pseudomonas were the only ones with high abundance. Bacillus 1 was the most abundant taxa recovered from surface sterilized tomato plant stems, in greenhouse and field experiments (Chapter 2 and 3 of this dissertation). Pseudomonas 1 was recovered in higher amount from tomato plants grown in the field. These genera are frequently isolated from agricultural crops (Miliute et al., 2015; Souza et al., 2013). Additionally, these genera have been isolated from tomato plants and exhibited beneficial activities towards various plant hosts, such as indole acetic acid production, nitrogen fixation and disease suppression (Duijff et al.,

1997; Nawangsih et al., 2011; Santoyo et al., 2016; Shahzad et al., 2017; Tamošiūnė et al., 2018; Tian et al., 2017).

Proteobacteria and Actinobacteria were the most abundant between tomato plant tissue and rhizosphere soil samples. Abundance of these phyla is common in plants, including tomato (Ding & Melcher, 2016; Dong et al., 2019; Rosenzweig et al., 2011;

Zhang et al., 2013). After assessing the bacterial diversity in the tomato rhizosphere

126 through culture-dependent and -independent approaches, Lee et al. 2016 found the majority of isolates and operational taxonomic units belonging to Proteobacteria and

Actinobacteria. Antoniou et al. 2017 evaluated the impact of compost on tomato plant growth by isolating bacteria from compost and the rhizosphere and found Proteobacteria and Actinobacteria highly enriched in the rhizosphere. In our study, taxa belonging to

Microbacterium, from the phylum Actinobacteria, was recovered across all treatments, plant tissues and sampling points. Microbacterium is predominant in various plants, including tomato, and has been studied for growth promotion in plants, including tomato

(Cordovez et al., 2018; Lin et al., 2012; Madhaiyan et al., 2010; Marquez-Santacruz et al., 2010; Zazou et al., 2016). Predominance of these taxonomic groups across soil and plant tissues might be due to horizontal and vertical transmission of microorganisms.

Seeds and roots could be colonized by bacteria in the soil, and eventually migrate to aboveground compartments such as stems and leaves (Compant et al., 2005; Delmotte et al., 2009).

As the active organic matter pool was higher in CGPPs, higher bacterial richness, and diversity in soil and plant tissue samples was observed in May and July. Increase in microbial diversity has been associated to the application of organic amendments

(Chaudhry et al., 2012; Sun et al., 2004). Enrichment in organic matter due to incorporation of chicken manure can increase microbial activity in soil (Gomez et al.,

2006). A study comparing an organic farming system with rich substrate in soil through the introduction of cattle farmyard manure, biological practices without the use synthetic

127 compounds and the presence of weed species, demonstrated higher microbial diversity compared to a conventional system (Lupatini et al., 2017). Due to availability of organic carbon and nitrogen, soils with higher amounts of C should favor copiotrophic bacteria, which thrive in environments with higher organic matter, as opposed to soils with lower amounts of C where oligotrophic bacteria, which thrive in environments with lower organic matter, should predominate (Ding et al., 2014; Hartmann et al., 2015). Soil and tomato plants grown in CGPPs could be harboring more copiotrophic bacteria, in comparison to POPs which had a lower active organic matter pool. Fierer et al. 2007 found a positive correlation of abundances of alpha Proteobacteria and Bacteroidetes with

C mineralization rates, hence, being considered copiotrophic bacteria. Similar to our results, these phyla were the most abundant in soil samples, regardless of chicken-grazing history. According to Ho et al. (2017), Acidobacteria and are more competitive under oligotrophic conditions. One of the least abundant phyla found in our study was Acidobacteria.

Although history of chicken grazing did not produce a significant shift in bacterial communities, it did affect relative abundance of individual taxa, as revealed by amplicon sequencing. A total of 65 ASVs were differentially abundant in stem and rhizosphere soil samples collected in July from CGPPs compared to POPs. Seven ASVs identified in rhizosphere soil in July belonged to the genus . This genus has been previously found as part of rhizospheric communities (Agnolucci et al., 2019; Jin et al.,

2020; Thokchom et al., 2017). Two ASVs from the Chitinophagaceae family were

128 differentially abundant in rhizosphere soil. These bacteria have been previously isolated from soil samples, and are known to encode several proteins predicted to have β- glucosidase activity, which is important in cellulose composition (Bailey et al., 2013).

Moreover, there was an increase of ASVs from the Rhizobiales order. Bacteria in the

Rhizobiales are associated to C-cycling in soil and N-cycling in leguminous plants (Fan et al., 2014). Five ASVs identified in rhizosphere soil and stem in July belong to the genus Sphingomonas. This genus has been found as a plant endophyte, including the culturing surveys in this dissertation, and has growth promoting effects in tomato plants

(Khan et al., 2014). It would be important to evaluate application rate of chicken manure and determine the extent of its impact on differentially abundant ASVs and their possible direct and indirect effects on the plant host.

Considering chicken manure is known to harbor human pathogens that can potentially contaminate fresh produce, it might be the source of high abundance of

Enterobacteriaceae in the endophytic community of all plant tissues (Chen & Jiang, 2014;

Omeira et al., 2006). Chicken manure had 82% Firmicutes followed by 17%

Proteobacteria, which is consistent with reports from previous studies (Gong et al., 2007;

Zhang et al., 2018). Additionally, chicken manure samples contained families such as

Lactobacillaceae (56%), Enterobacteriaceae (15%) and Enterococcaceae (10%) which have been previously found in chicken’s intestinal and litter microbiota (Cressman et al.,

2010). According to amplicons sequencing data, Enterobacteriaceae was found in high abundance in all plant tissues, dominating tomato flesh (42%) and tomato seed (37%),

129 regardless of chicken grazing history. Bacteria in this family are best known for their pathogenic effect on humans, as well as their presence in fresh and processed tomato (Al-

Kharousi et al., 2016; Eribo & Ashenafi, 2003). Recovered isolates from taxa belonging to the Enterobacteriaceae family was present in both treatments, sampling months, and plant tissues. Increased abundance of Enterobacteriaceae can also be attributed to elevated temperature and humidity, as seen by Allard et al. (2020) in epiphytic communities on tomato fruit. There are bacteria within Enterobacteriaceae include plant endophytes with plant growth promoting activity in tomato (Bendaha & Belaouni, 2019;

Gupta et al., 2020; Mayak et al., 2001). Examples of taxa belonging to

Enterobacteriaceae recovered in our study include Pantoea and Enterobacter. Pantoea was recovered across all treatments, plant tissues and sampling points. Pantoea has studied for its nitrogen fixation and plant growth-promoting capabilities in tomato

(Dursun et al., 2010; Enya et al., 2007; Enya et al., 2007; Sharon et al., 2016; Walpola &

Yoon, 2013; Walterson & Stavrinides, 2015). On the other hand, Pantoea also known as a plant and human pathogen (Coutinho & Venter, 2009; Cruz et al., 2007). Enterobacter has been studied as a beneficial plant endophyte in tomato and banana (Bendaha &

Belaouni, 2019; Macedo-Raygoza et al., 2019; Telias et al., 2011). On the other hand,

Enterobacter had been studied as an opportunistic human pathogen, as well as exhibiting multiple antibiotic resistance (Mishra et al., 2017). Further research is needed to identify possible harmful bacterial endophytes in soil and tomatoes plant tissues.

130

Implementing chicken manure in agricultural practices has shown to reduce plant disease incidence (Pan et al., 2017), therefore isolates recovered from tomato plants grown in CGPPs have potentially antagonistic properties towards plant pathogens. In this study, a variety of bacteria were isolated, some of which belong to genera with previously described antagonistic properties, such as Bacillus, Pseudomonas, Paenibacillus and

Stenotrophomonas (Massart et al., 2015). Nandhini et al. (2012) recovered endophytic bacteria from surface sterilized root, stem, leaves and fruits of healthy tomato plants.

Through a screening process, only 50% of the bacterial isolates showed antagonistic activity, including Bacillus and Pseudomonas. Conn and Lazarovits (1999) observed a reduction in verticillium wilt, potato scab incidence and plant-parasitic nematodes in potato crops in different field sites incorporated with chicken manure. A study conducted by Zhang et al. (2020) incorporated a mixture of biochar with fresh chicken manure and observed a significant decrease of Fusarium spp and Phytophthora spp. They used high- throughput gene sequencing to monitor changes in the soil’s bacterial and fungal communities and found a significant decrease of the diversity of these communities.

Consequently, the authors observed greater relative abundance of Actinobacteria and the genera Pseudomonas, Bacillus and Chaetomium, which are known to have antagonistic properties.

Sample origin and sampling month were significant drivers in clustering of endophytic bacterial composition in stem, root and rhizosphere samples. In this study, we were able to observe how bacterial communities inside the plant host are organ specific.

131

A study by Allard et al. (2016) assessed rhizosphere and phyllosphere bacterial communities of tomato plants that were grown in soils with different organic amendments, including fresh poultry litter, and had similar findings. However, Allard et al. (2016) limited their sampling to one time point and were not able to address differences during plant development. In a different study, Gao et al. (2015) observed higher soil microbial diversity in cucumber rhizosphere during earlier plant stages compared to the following months. Richness and diversity of bacterial communities in soil and plant tissue samples varied in response to sampling month. In May and July, there was a trend of higher bacterial richness, evenness and phylogenetic diversity for soil and plant tissue samples obtained from CGPPs. In September, however, richness, evenness and phylogenetic diversity was lower for all samples obtained from CGPPs, with the exception of seed and leaf. According to plant development stage, leaf, stem, root and rhizosphere-associated microbial populations can vary in abundance community composition (Carper et al., 2018; Chaparro et al., 2012; Copeland et al., 2015; Wagner et al., 2016). Changes varying by plant development stage could be attributed to shifts in the host plant’s nutritional needs (Hacquard et al., 2015; Panke-Buisse et al., 2015).

Considering its significantly higher microbial activity, soil in CGPPs might harbor more bacteria with functional diversity, such as plant growth promotion, affecting the tomato plants initial selection during early plant development stages. Edwards et al. (2018) found the root microbiota of rice plants to be highly dynamic during the vegetative phase of plant growth and followed by stabilization of its composition for the remainder of the life cycle.

132

Variations in amplicon sequencing data within sample origin can be improved through better standardization. Plant microbiome surveys tend to use high-throughput amplicon sequencing. This method, coupled with well-designed sampling strategies, unveils the composition, organization, and spatial distribution of microbial communities in a given sample (Knief et al., 2012). Amplicon sequencing can be used for targeting single groups of microbes, rare organisms, as well as functional genes (Herbold et al.,

2015). However, its sensitivity makes it more prone to contamination and the addition of positive and negative controls is necessary (Glassing et al., 2016). In this study, a positive control was not included with the samples sent to sequencing. Including positive controls is important to ensure there are no false positive or false negative results. The lack of positive controls is common in microbiome research. Hornung et al. (2019) reviewed all publications from the 2018 issues of Microbiome and the ISME journal and found that from the 265 publications utilizing high-throughput community sequencing of any type, only 30% reported using any type of negative control, and only 10% reported using a positive control. It is difficult to compare results of cultivation-independent approaches due to the use of different sampling protocols, primers, and sequencing pipelines. When profiling communities by targeting the 16S rRNA gene using NGS techniques, a recurrent problem is the contamination of samples with undesired sequences, such as chloroplast or mitochondrial DNA (Lutz et al., 2011). Some studies in plant microbiomes employ the use of primers that minimize contamination (Ghyselinck et al., 2013). In this study, we used primer pair 799F/1391R which have been found to amplify the

133 hypervariable regions V5, V6 and V7 of the 16S gene and exclude chloroplast and mitochondrial DNA (Beckers et al., 2016; Bodenhausen et al., 2013; Horton et al., 2015).

Due to differences in sequencing depth for samples from different origins, normalization had to be performed. Although common in microbiome research, rarefying counts was not used to normalize data. Rarefying counts is known to produce a high rate of false positives in tests for species that are differentially abundant across sample classes, as well as discards samples (McMurdie & Holmes, 2014). CSS is the most appropriate normalization method to be able to determine differences in rare ASVs between treatments, sampling months and plots.

Of all samples processed, rhizosphere soil and root samples had the highest sequencing depth, and the highest richness and diversity. This is consistent with a study by Ottesen et al. (2013), where the authors saw a higher number of OTU's observed for roots compared to the rest of the tomato plant tissues. Bergna et al. (2018) also recorded higher alpha diversity measurements for rhizosphere soil and root and lower in tomato seed. Although significant differences were found between plots, leaf samples recorded an unexpectedly high richness and diversity. Therefore, higher richness and diversity in bacterial communities from leaf samples in this study might be due to inefficiency in the surface sterilization process, which in turn can affect the detection of epiphytes, and therefore increasing the diversity of bacteria recovered in the leaf sample. Although not common in plant microbiome studies, we used chloroplast and mitochondria excluding primers, which increases the coverage of bacterial sequences, ultimately affecting

134 diversity (Lucaciu et al., 2019). Selection of primers can be a source of bias, therefore, comparisons between studies must take this into account.

With this study, we were able to characterize the endophytic bacterial microbiome of tomato plants grown in CGPPs. Through the use of culture-dependent and culture- independent techniques, we were able to determine the richness and abundance of bacterial communities in soil, tomato plant tissues and chicken manure. Additionally, we were able to assess the impact of management strategy, sample origin and sampling point on these bacterial communities, as well as soil health and plant growth. Our results showed that the sample origin, whether it is soil or above- and belowground plant compartments has a larger influence on bacterial richness and abundance, than agricultural management strategy. Our findings suggest stronger contribution of agricultural management practices during early plant stages on endophytic microbiome, as opposed to later on in the host lifecycle. Improvements in standardization are important preparing samples for sequencing to avoid any biases and incorrect assumptions when analyzing the data. Additionally, maintaining consistency with samples size are essential when attempting to interpret data with sufficient statistical support. Further directions should focus on identification and elucidation of transmission of possible plant growth promoting bacteria, as well as human pathogens, found in crops grown in this type of agricultural management strategy. Moreover, the influence of stage in plant lifecycle on bacterial community composition under specific management strategies remains to be explored.

135

References

Agnolucci, M., Palla, M., Cristani, C., Cavallo, N., Giovannetti, M., De Angelis, M., Gobbetti, M., & Minervini, F. (2019). Beneficial Plant Microorganisms Affect the Endophytic Bacterial Communities of Durum Wheat Roots as Detected by Different Molecular Approaches. Frontiers in Microbiology, 10. https://doi.org/10.3389/fmicb.2019.02500

Alaoui, A., Rogger, M., Peth, S., & Blöschl, G. (2018). Does soil compaction increase floods? A review. Journal of Hydrology, 557, 631–642. https://doi.org/10.1016/j.jhydrol.2017.12.052

Al-Kharousi, Z. S., Guizani, N., Al-Sadi, A. M., Al-Bulushi, I. M., & Shaharoona, B. (2016). Hiding in Fresh Fruits and Vegetables: Opportunistic Pathogens May Cross Geographical Barriers [Research Article]. International Journal of Microbiology; Hindawi. https://doi.org/10.1155/2016/4292417

Allard, S. M., Ottesen, A. R., & Micallef, S. A. (2020). Rain induces temporary shifts in epiphytic bacterial communities of cucumber and tomato fruit. Scientific Reports, 10(1), 1765. https://doi.org/10.1038/s41598-020-58671-7

Allard, S. M., Walsh, C. S., Wallis, A. E., Ottesen, A. R., Brown, E. W., & Micallef, S. A. (2016). Solanum lycopersicum (tomato) hosts robust phyllosphere and rhizosphere bacterial communities when grown in soil amended with various organic and synthetic fertilizers. Science of The Total Environment, 573, 555– 563. https://doi.org/10.1016/j.scitotenv.2016.08.157

Altschul, S. F., Gish, W., Miller, W., Myers, E. W., & Lipman, D. J. (1990). Basic local alignment search tool. Journal of Molecular Biology, 215(3), 403–410. https://doi.org/10.1016/S0022-2836(05)80360-2

Andersen, K. S., Kirkegaard, R. H., Karst, S. M., & Albertsen, M. (2018). ampvis2: An R package to analyse and visualise 16S rRNA amplicon data. BioRxiv, 299537. https://doi.org/10.1101/299537

Anderson, M. J. (2017). Permutational Multivariate Analysis of Variance (PERMANOVA). In Wiley StatsRef: Statistics Reference Online (pp. 1–15). American Cancer Society. https://doi.org/10.1002/9781118445112.stat07841

Antoniou, A., Tsolakidou, M.-D., Stringlis, I. A., & Pantelides, I. S. (2017). Rhizosphere Microbiome Recruited from a Suppressive Compost Improves

136

Plant Fitness and Increases Protection against Vascular Wilt Pathogens of Tomato. Frontiers in Plant Science, 8. https://doi.org/10.3389/fpls.2017.02022

Araji, A. A., Abdo, Z. O., & Joyce, P. (2001). Efficient use of animal manure on cropland—Economic analysis. Bioresource Technology, 79(2), 179–191. https://doi.org/10.1016/s0960-8524(01)00042-6

Bailey, V. L., Fansler, S. J., Stegen, J. C., & McCue, L. A. (2013). Linking microbial community structure to β -glucosidic function in soil aggregates. The ISME Journal, 7(10), 2044–2053. https://doi.org/10.1038/ismej.2013.87

Beckers, B., Op De Beeck, M., Thijs, S., Truyens, S., Weyens, N., Boerjan, W., & Vangronsveld, J. (2016). Performance of 16s rDNA Primer Pairs in the Study of Rhizosphere and Endosphere Bacterial Microbiomes in Metabarcoding Studies. Frontiers in Microbiology, 7. https://doi.org/10.3389/fmicb.2016.00650

Bendaha, M. E. A., & Belaouni, H. A. (2019). Tomato growth and resistance promotion by Enterobacter hormaechei subsp. Steigerwaltii EB8D. Archives of Phytopathology and Plant Protection, 52(3–4), 318–332. https://doi.org/10.1080/03235408.2019.1620511

Bergna, A., Cernava, T., Rändler, M., Grosch, R., Zachow, C., & Berg, G. (2018). Tomato Seeds Preferably Transmit Plant Beneficial Endophytes. Phytobiomes Journal, 2(4), 183–193. https://doi.org/10.1094/PBIOMES-06-18-0029-R

Bergougnoux, V. (2014). The history of tomato: From domestication to biopharming. Biotechnology Advances, 32(1), 170–189. https://doi.org/10.1016/j.biotechadv.2013.11.003

Bishop, J. P. (1995). Chickens: Improving small-scale production. Echo Technical Note, 10.

Bodenhausen, N., Horton, M. W., & Bergelson, J. (2013). Bacterial Communities Associated with the Leaves and the Roots of Arabidopsis thaliana. PLoS ONE, 8(2), e56329. https://doi.org/10.1371/journal.pone.0056329

Callahan, B. J., McMurdie, P. J., Rosen, M. J., Han, A. W., Johnson, A. J. A., & Holmes, S. P. (2016). DADA2: High-resolution sample inference from Illumina amplicon data. Nature Methods, 13(7), 581–583. https://doi.org/10.1038/nmeth.3869

Carper, D. L., Carrell, A. A., Kueppers, L. M., & Frank, A. C. (2018). Bacterial endophyte communities in Pinus flexilis are structured by host age, tissue type,

137

and environmental factors. Plant and Soil; Dordrecht, 428(1–2), 335–352. http://dx.doi.org.proxy.lib.ohio-state.edu/10.1007/s11104-018-3682-x

CFAES. (2020). College of Food, Agricultural, and Environmental Sciences Weather System: Wooster Station. CFAES Weather System, The Ohio State University. https://www.oardc.ohio-state.edu/weather1/

Chao, A., Chazdon, R. L., Colwell, R. K., & Shen, T.-J. (2006). Abundance-based similarity indices and their estimation when there are unseen species in samples. Biometrics, 62(2), 361–371. https://doi.org/10.1111/j.1541-0420.2005.00489.x

Chaparro, J. M., Sheflin, A. M., Manter, D. K., & Vivanco, J. M. (2012). Manipulating the soil microbiome to increase soil health and plant fertility. Biology and Fertility of Soils, 48(5), 489–499. https://doi.org/10.1007/s00374-012-0691-4

Chaudhry, V., Rehman, A., Mishra, A., Chauhan, P. S., & Nautiyal, C. S. (2012). Changes in Bacterial Community Structure of Agricultural Land Due to Long- Term Organic and Chemical Amendments. Microbial Ecology, 64(2), 450–460. https://doi.org/10.1007/s00248-012-0025-y

Chelius, M. K., & Triplett, E. W. (2001). The Diversity of Archaea and Bacteria in Association with the Roots of Zea mays L. Microbial Ecology, 41(3), 252–263. https://doi.org/10.1007/s002480000087

Chellemi, D. O., Rosskopf, E. N., & Kokalis-Burelle, N. (2013). The Effect of Transitional Organic Production Practices on Soilborne Pests of Tomato in a Simulated Microplot Study. Phytopathology, 103(8), 792–801. https://doi.org/10.1094/PHYTO-09-12-0243-R

Chen, Z., & Jiang, X. (2014). Microbiological Safety of Chicken Litter or Chicken Litter-Based Organic Fertilizers: A Review. Agriculture, 4(1), 1–29. https://doi.org/10.3390/agriculture4010001

Cheng, Z., Lei, S., Li, Y., Huang, W., Ma, R., Xiong, J., Zhang, T., Jin, L., Haq, H. ul, Xu, X., & Tian, B. (2020). Revealing the Variation and Stability of Bacterial Communities in Tomato Rhizosphere Microbiota. Microorganisms, 8(2). https://doi.org/10.3390/microorganisms8020170

Ch’ng, H. Y., Ahmed, O. H., Kassim, S., & Majid, N. M. A. (2013). Co-composting of pineapple leaves and chicken manure slurry. International Journal Of Recycling of Organic Waste in Agriculture, 2(1), 23. https://doi.org/10.1186/2251-7715-2- 23

138

Compant, S., Reiter, B., Sessitsch, A., Nowak, J., Clément, C., & Barka, E. A. (2005). Endophytic Colonization of Vitis vinifera L. by Plant Growth-Promoting Bacterium Burkholderia sp. Strain PsJN. Applied and Environmental Microbiology, 71(4), 1685–1693. https://doi.org/10.1128/AEM.71.4.1685- 1693.2005

Conn, K. L., & Lazarovits, G. (1999). Impact of animal manures on verticillium wilt, potato scab, and soil microbial populations. Canadian Journal of Plant Pathology, 21(1), 81–92. https://doi.org/10.1080/07060661.1999.10600089

Copeland, J. K., Yuan, L., Layeghifard, M., Wang, P. W., & Guttman, D. S. (2015). Seasonal Community Succession of the Phyllosphere Microbiome. Molecular Plant-Microbe Interactions®, 28(3), 274–285. https://doi.org/10.1094/MPMI- 10-14-0331-FI

Cordovez, V., Schop, S., Hordijk, K., Boulois, H. D. de, Coppens, F., Hanssen, I., Raaijmakers, J. M., & Carrión, V. J. (2018). Priming of Plant Growth Promotion by Volatiles of Root-Associated Microbacterium spp. Applied and Environmental Microbiology, 84(22). https://doi.org/10.1128/AEM.01865-18

Coutinho, T. A., & Venter, S. N. (2009). Pantoea ananatis: An unconventional plant pathogen. Molecular Plant Pathology, 10(3), 325–335. https://doi.org/10.1111/j.1364-3703.2009.00542.x

Cressman, M. D., Yu, Z., Nelson, M. C., Moeller, S. J., Lilburn, M. S., & Zerby, H. N. (2010). Interrelations between the Microbiotas in the Litter and in the Intestines of Commercial Broiler Chickens. Applied and Environmental Microbiology, 76(19), 6572–6582. https://doi.org/10.1128/AEM.00180-10

Cruz, A. T., Cazacu, A. C., & Allen, C. H. (2007). Pantoea agglomerans, a Plant Pathogen Causing Human Disease. Journal of Clinical Microbiology, 45(6), 1989–1992. https://doi.org/10.1128/JCM.00632-07

Culman, S. W., Snapp, S. S., Freeman, M. A., Schipanski, M. E., Beniston, J., Lal, R., Drinkwater, L. E., Franzluebbers, A. J., Glover, J. D., Grandy, A. S., Lee, J., Six, J., Maul, J. E., Mirksy, S. B., Spargo, J. T., & Wander, M. M. (2012). Permanganate Oxidizable Carbon Reflects a Processed Soil Fraction that is Sensitive to Management. Soil Science Society of America Journal, 76(2), 494– 504. https://doi.org/10.2136/sssaj2011.0286

Davis, N. M., Proctor, D. M., Holmes, S. P., Relman, D. A., & Callahan, B. J. (2018). Simple statistical identification and removal of contaminant sequences in marker-gene and metagenomics data. Microbiome, 6(1), 226. https://doi.org/10.1186/s40168-018-0605-2 139

de Mendiburu, F. (2020). agricolae: Statistical Procedures for Agricultural Research. (Version R package version 1.3-2) [Computer software]. https://CRAN.R- project.org/package=agricolae

DEFRA. (2001). The welfare of hens in free range systems. Department for Environment, Food & Rural Affairs, U.K.

Delmotte, N., Knief, C., Chaffron, S., Innerebner, G., Roschitzki, B., Schlapbach, R., Mering, C. von, & Vorholt, J. A. (2009). Community proteogenomics reveals insights into the physiology of phyllosphere bacteria. Proceedings of the National Academy of Sciences, 106(38), 16428–16433. https://doi.org/10.1073/pnas.0905240106

Demir, K., Sahin, O., Kadioglu, Y. K., Pilbeam, D. J., & Gunes, A. (2010). Essential and non-essential element composition of tomato plants fertilized with poultry manure. Scientia Horticulturae, 127(1), 16–22. https://doi.org/10.1016/j.scienta.2010.08.009

Devendra, C., & Thomas, D. (2002). Crop–animal interactions in mixed farming systems in Asia. Agricultural Systems, 71(1–2), 27–40. https://doi.org/10.1016/S0308-521X(01)00034-8

Ding, G.-C., Radl, V., Schloter-Hai, B., Jechalke, S., Heuer, H., Smalla, K., & Schloter, M. (2014). Dynamics of Soil Bacterial Communities in Response to Repeated Application of Manure Containing Sulfadiazine. PLOS ONE, 9(3), e92958. https://doi.org/10.1371/journal.pone.0092958

Ding, T., & Melcher, U. (2016). Influences of Plant Species, Season and Location on Leaf Endophytic Bacterial Communities of Non-Cultivated Plants. PLoS ONE, 11(3). https://doi.org/10.1371/journal.pone.0150895

Dong, C.-J., Wang, L.-L., Li, Q., & Shang, Q.-M. (2019). Bacterial communities in the rhizosphere, phyllosphere and endosphere of tomato plants. PLoS ONE, 14(11). https://doi.org/10.1371/journal.pone.0223847

Duijff, B. J., Gianinazzi‐Pearson, V., & Lemanceau, P. (1997). Involvement of the outer membrane lipopolysaccharides in the endophytic colonization of tomato roots by biocontrol Pseudomonas fluorescens strain WCS417r. New Phytologist, 135(2), 325–334. https://doi.org/10.1046/j.1469-8137.1997.00646.x

Dursun, A., Ekinci, M., & Dönmez, M. F. (2010). Effects of foliar application of plant growth promoting bacterium on chemical contents, yield and growth of tomato

140

(Lycopersicon esculentum L.) and cucumber (Cucumis sativus L.). Pakistan Journal of Botany, 42, 3349–3356.

Edwards, J. A., Santos-Medellín, C. M., Liechty, Z. S., Nguyen, B., Lurie, E., Eason, S., Phillips, G., & Sundaresan, V. (2018). Compositional shifts in root-associated bacterial and archaeal microbiota track the plant life cycle in field-grown rice. PLOS Biology, 16(2), e2003862. https://doi.org/10.1371/journal.pbio.2003862

Enya, J., Koitabashi, M., Shinohara, H., Yoshida, S., Tsukiboshi, T., Negishi, H., Suyama, K., & Tsushima, S. (2007). Phylogenetic Diversities of Dominant Culturable Bacillus, Pseudomonas and Pantoea Species on Tomato Leaves and Their Possibility as Biological Control Agents. Journal of Phytopathology, 155(7–8), 446–453. https://doi.org/10.1111/j.1439-0434.2007.01256.x

Enya, J., Shinohara, H., Yoshida, S., Tsukiboshi, T., Negishi, H., Suyama, K., & Tsushima, S. (2007). Culturable Leaf-Associated Bacteria on Tomato Plants and Their Potential as Biological Control Agents. Microbial Ecology, 53(4), 524– 536. https://doi.org/10.1007/s00248-006-9085-1

Eribo, B., & Ashenafi, M. (2003). Behavior of Escherichia coli O157:H7 in tomato and processed tomato products. Food Research International, 36(8), 823–830. https://doi.org/10.1016/S0963-9969(03)00077-2

Faith, D. P. (2018). Phylogenetic Diversity and Conservation Evaluation: Perspectives on Multiple Values, Indices, and Scales of Application. In R. A. Scherson & D. P. Faith (Eds.), Phylogenetic Diversity (pp. 1–26). Springer International Publishing. https://doi.org/10.1007/978-3-319-93145-6_1

Fan, F., Yin, C., Tang, Y., Li, Z., Song, A., Wakelin, S. A., Zou, J., & Liang, Y. (2014). Probing potential microbial coupling of carbon and nitrogen cycling during decomposition of maize residue by 13C-DNA-SIP. Soil Biology and Biochemistry, 70, 12–21. https://doi.org/10.1016/j.soilbio.2013.12.002

FAO. (2018). FAOSTAT: Food and agriculture data. Food and Agriculture Organization of the United Nations. http://www.fao.org/faostat/en/

Fierer, N., Bradford, M. A., & Jackson, R. B. (2007). Toward an ecological classification of soil bacteria. Ecology, 88(6), 1354–1364. https://doi.org/10.1890/05-1839

Fox, J., & Weisberg, S. (2019). An {R} Companion to Applied Regression (Third). Sage. https://socialsciences.mcmaster.ca/jfox/Books/Companion/

141

Franzluebbers, A. J., Haney, R. L., Honeycutt, C. W., Schomberg, H. H., & Hons, F. M. (2000). Flush of Carbon Dioxide Following Rewetting of Dried Soil Relates to Active Organic Pools. Soil Science Society of America Journal, 64(2), 613–623. https://doi.org/10.2136/sssaj2000.642613x

Gao, Y., Tian, Y., Liang, X., & Gao, L. (2015). Effects of single-root-grafting, double- root-grafting and compost application on microbial properties of rhizosphere soils in Chinese protected cucumber (Cucumis sativus L.) production systems. Scientia Horticulturae, 186, 190–200. https://doi.org/10.1016/j.scienta.2015.02.026

Genomics Workbench version 20.0. (2020). https://digitalinsights.qiagen.com

Gerberastraat, B. V., & Aalsmeer, R. A. (2001). 4Peaks. Nucleobytes. https://nucleobytes.com/4peaks/index.html

Ghyselinck, J., Pfeiffer, S., Heylen, K., Sessitsch, A., & De Vos, P. (2013). The Effect of Primer Choice and Short Read Sequences on the Outcome of 16S rRNA Gene Based Diversity Studies. PLoS ONE, 8(8), e71360. https://doi.org/10.1371/journal.pone.0071360

Glassing, A., Dowd, S. E., Galandiuk, S., Davis, B., & Chiodini, R. J. (2016). Inherent bacterial DNA contamination of extraction and sequencing reagents may affect interpretation of microbiota in low bacterial biomass samples. Gut Pathogens, 8(1), 24. https://doi.org/10.1186/s13099-016-0103-7

Gomez, E., Ferreras, L., & Toresani, S. (2006). Soil bacterial functional diversity as influenced by organic amendment application. Bioresource Technology, 97(13), 1484–1489. https://doi.org/10.1016/j.biortech.2005.06.021

Gong, J., Si, W., Forster, R. J., Huang, R., Yu, H., Yin, Y., Yang, C., & Han, Y. (2007). 16S rRNA gene-based analysis of mucosa-associated bacterial community and phylogeny in the chicken gastrointestinal tracts: From crops to ceca. FEMS Microbiology Ecology, 59(1), 147–157. https://doi.org/10.1111/j.1574- 6941.2006.00193.x

Gupta, P., Kumar, V., Usmani, Z., Rani, R., Chandra, A., & Gupta, V. K. (2020). Implications of plant growth promoting Klebsiella sp. CPSB4 and Enterobacter sp. CPSB49 in luxuriant growth of tomato plants under chromium stress. Chemosphere, 240, 124944. https://doi.org/10.1016/j.chemosphere.2019.124944

Hacquard, S., Garrido-Oter, R., González, A., Spaepen, S., Ackermann, G., Lebeis, S., McHardy, A. C., Dangl, J. L., Knight, R., Ley, R., & Schulze-Lefert, P. (2015).

142

Microbiota and Host Nutrition across Plant and Animal Kingdoms. Cell Host & Microbe, 17(5), 603–616. https://doi.org/10.1016/j.chom.2015.04.009

Hartman, K., van der Heijden, M. G. A., Wittwer, R. A., Banerjee, S., Walser, J.-C., & Schlaeppi, K. (2018). Cropping practices manipulate abundance patterns of root and soil microbiome members paving the way to smart farming. Microbiome, 6(1), 14. https://doi.org/10.1186/s40168-017-0389-9

Hartmann, M., Frey, B., Mayer, J., Mäder, P., & Widmer, F. (2015). Distinct soil microbial diversity under long-term organic and conventional farming. The ISME Journal, 9(5), 1177–1194. https://doi.org/10.1038/ismej.2014.210

Herbold, C. W., Pelikan, C., Kuzyk, O., Hausmann, B., Angel, R., Berry, D., & Loy, A. (2015). A flexible and economical barcoding approach for highly multiplexed amplicon sequencing of diverse target genes. Frontiers in Microbiology, 6. https://doi.org/10.3389/fmicb.2015.00731

Hilimire, K., Gliessman, S., & Muramoto, J. (2013). Soil fertility and crop growth under poultry/crop integration. Renewable Agriculture and Food Systems, 28, 173– 182. https://doi.org/10.1017/S174217051200021X

Ho, A., Di Lonardo, D. P., & Bodelier, P. L. E. (2017). Revisiting life strategy concepts in environmental microbial ecology. FEMS Microbiology Ecology, 93(3). https://doi.org/10.1093/femsec/fix006

Hole, D. G., Perkins, A. J., Wilson, J. D., Alexander, I. H., Grice, P. V., & Evans, A. D. (2005). Does organic farming benefit biodiversity? Biological Conservation, 122(1), 113–130. https://doi.org/10.1016/j.biocon.2004.07.018

Hornung, B. V. H., Zwittink, R. D., & Kuijper, E. J. (2019). Issues and current standards of controls in microbiome research. FEMS Microbiology Ecology, 95(5). https://doi.org/10.1093/femsec/fiz045

Horton, M. W., Bodenhausen, N., Beilsmith, K., Meng, D., Subramanian, S., Vetter, M. M., Vilhjálmsson, B. J., Gordon, J. I., & Bergelson, J. (2015). Genome-wide association study of Arabidopsis thaliana’s leaf microbial community. Nature Communications, 5(1), 1–7.

Hurisso, T. T., Moebius‐Clune, D. J., Culman, S. W., Moebius‐Clune, B. N., Thies, J. E., & Es, H. M. van. (2018). Soil Protein as a Rapid Soil Health Indicator of Potentially Available Organic Nitrogen. Agricultural & Environmental Letters, 3(1), 180006. https://doi.org/10.2134/ael2018.02.0006

143

Jin, X., Shi, Y., Wu, F., Pan, K., Zhou, X., Jin, X., Shi, Y., Wu, F., Pan, K., & Zhou, X. (2020). Intercropping of wheat changed cucumber rhizosphere bacterial community composition and inhibited cucumber Fusarium wilt disease. Scientia Agricola, 77(5). https://doi.org/10.1590/1678-992x-2019-0005

Khan, A. L., Waqas, M., Kang, S.-M., Al-Harrasi, A., Hussain, J., Al-Rawahi, A., Al- Khiziri, S., Ullah, I., Ali, L., Jung, H.-Y., & Lee, I.-J. (2014). Bacterial endophyte Sphingomonas sp. LK11 produces gibberellins and IAA and promotes tomato plant growth. Journal of Microbiology, 52(8), 689–695. https://doi.org/10.1007/s12275-014-4002-7

Knief, C., Delmotte, N., Chaffron, S., Stark, M., Innerebner, G., Wassmann, R., Mering, C. von, & Vorholt, J. A. (2012). Metaproteogenomic analysis of microbial communities in the phyllosphere and rhizosphere of rice. The ISME Journal, 6(7), 1378–1390. https://doi.org/10.1038/ismej.2011.192

Kwak, M.-J., Kong, H. G., Choi, K., Kwon, S.-K., Song, J. Y., Lee, J., Lee, P. A., Choi, S. Y., Seo, M., Lee, H. J., Jung, E. J., Park, H., Roy, N., Kim, H., Lee, M. M., Rubin, E. M., Lee, S.-W., & Kim, J. F. (2018). Rhizosphere microbiome structure alters to enable wilt resistance in tomato. Nature Biotechnology, 36(11), 1100–1109. https://doi.org/10.1038/nbt.4232

Kyakuwaire, M., Olupot, G., Amoding, A., Nkedi-Kizza, P., & Ateenyi Basamba, T. (2019). How Safe is Chicken Litter for Land Application as an Organic Fertilizer?: A Review. International Journal of Environmental Research and Public Health, 16(19). https://doi.org/10.3390/ijerph16193521

Lai, L., Zhao, X., Jiang, L., Wang, Y., Luo, L., Zheng, Y., Chen, X., & Rimmington, G. M. (2012). Soil Respiration in Different Agricultural and Natural Ecosystems in an Arid Region. PLOS ONE, 7(10), e48011. https://doi.org/10.1371/journal.pone.0048011

Lee, S. A., Park, J., Chu, B., Kim, J. M., Joa, J.-H., Sang, M. K., Song, J., & Weon, H.- Y. (2016). Comparative analysis of bacterial diversity in the rhizosphere of tomato by culture-dependent and -independent approaches. Journal of Microbiology, 54(12), 823–831. https://doi.org/10.1007/s12275-016-6410-3

Lenth, R. V. (2016). Least-Squares Means: The R Package lsmeans. Journal of Statistical Software, 69(1), 1–33. https://doi.org/10.18637/jss.v069.i01

Lin, L., Guo, W., Xing, Y., Zhang, X., Li, Z., Hu, C., Li, S., Li, Y., & An, Q. (2012). The actinobacterium Microbacterium sp. 16SH accepts pBBR1-based pPROBE vectors, forms biofilms, invades roots, and fixes N2 associated with

144

micropropagated sugarcane plants. Applied Microbiology and Biotechnology, 93(3), 1185–1195. https://doi.org/10.1007/s00253-011-3618-3

López, S., Pastorino, G., Franco, M., Medina, R., Lucentini, C., & Balatti, P. (2018). Microbial Endophytes that Live within the Seeds of Two Tomato Hybrids Cultivated in Argentina. Agronomy, 8, 136. https://doi.org/10.3390/agronomy8080136

Love, M. I., Huber, W., & Anders, S. (2014). Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biology, 15(12), 550. https://doi.org/10.1186/s13059-014-0550-8

Lozupone, C., & Knight, R. (2005). UniFrac: A New Phylogenetic Method for Comparing Microbial Communities. Applied and Environmental Microbiology, 71(12), 8228–8235. https://doi.org/10.1128/AEM.71.12.8228-8235.2005

Lucaciu, R., Pelikan, C., Gerner, S. M., Zioutis, C., Köstlbacher, S., Marx, H., Herbold, C. W., Schmidt, H., & Rattei, T. (2019). A Bioinformatics Guide to Plant Microbiome Analysis. Frontiers in Plant Science, 10. https://doi.org/10.3389/fpls.2019.01313

Lupatini, M., Korthals, G. W., de Hollander, M., Janssens, T. K. S., & Kuramae, E. E. (2017). Soil Microbiome Is More Heterogeneous in Organic Than in Conventional Farming System. Frontiers in Microbiology, 7. https://doi.org/10.3389/fmicb.2016.02064

Lutz, K. A., Wang, W., Zdepski, A., & Michael, T. P. (2011). Isolation and analysis of high quality nuclear DNA with reduced organellar DNA for plant genome sequencing and resequencing. BMC Biotechnology, 11(1), 54. https://doi.org/10.1186/1472-6750-11-54

Macedo-Raygoza, G. M., Valdez-Salas, B., Prado, F. M., Prieto, K. R., Yamaguchi, L. F., Kato, M. J., Canto-Canché, B. B., Carrillo-Beltrán, M., Di Mascio, P., White, J. F., & Beltrán-García, M. J. (2019). Enterobacter cloacae, an Endophyte That Establishes a Nutrient-Transfer Symbiosis With Banana Plants and Protects Against the Black Sigatoka Pathogen. Frontiers in Microbiology, 10. https://doi.org/10.3389/fmicb.2019.00804

Madhaiyan, M., Poonguzhali, S., Lee, J.-S., Lee, K.-C., Saravanan, V. S., & Santhanakrishnan, P. (2010). Microbacterium azadirachtae sp. Nov., a plant- growth-promoting actinobacterium isolated from the rhizoplane of neem seedlings. International Journal of Systematic and Evolutionary Microbiology, 60(7), 1687–1692. https://doi.org/10.1099/ijs.0.015800-0

145

Marquez-Santacruz, H. A., Hernandez-Leon, R., Orozco-Mosqueda, M. C., Velazquez- Sepulveda, I., & Santoyo, G. (2010). Diversity of bacterial endophytes in roots of Mexican husk tomato plants (Physalis ixocarpa) and their detection in the rhizosphere. Genetics and Molecular Research, 9(4), 2372–2380. https://doi.org/10.4238/vol9-4gmr921

Martin, M. (2011). Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet.Journal, 17(1), 10–12. https://doi.org/10.14806/ej.17.1.200

Maru, A., Haruna, O. A., & Charles Primus, W. (2015). Coapplication of Chicken Litter Biochar and Urea Only to Improve Nutrients Use Efficiency and Yield of Oryza sativa L. Cultivation on a Tropical Acid Soil [Research Article]. The Scientific World Journal; Hindawi. https://doi.org/10.1155/2015/943853

Massart, S., Martinez-Medina, M., & Jijakli, M. H. (2015). Biological control in the microbiome era: Challenges and opportunities. Biological Control, 89, 98–108. https://doi.org/10.1016/j.biocontrol.2015.06.003

Mayak, S., Tirosh, T., & Glick, B. R. (2001). Stimulation of the Growth of Tomato, Pepper and Mung Bean Plants by the Plant Growth- Promoting Bacterium Enterobacter cloacae CAL3. Biological Agriculture & Horticulture, 19(3), 261– 274. https://doi.org/10.1080/01448765.2001.9754929

McCarty, G. W., Lyssenko, N. N., & Starr, J. L. (1998). Short-term Changes in Soil Carbon and Nitrogen Pools during Tillage Management Transition. Soil Science Society of America Journal, 62(6), 1564–1571. https://doi.org/10.2136/sssaj1998.03615995006200060013x

McKnight, D. T., Huerlimann, R., Bower, D. S., Schwarzkopf, L., Alford, R. A., & Zenger, K. R. (2019). Methods for normalizing microbiome data: An ecological perspective. Methods in Ecology and Evolution, 10(3), 389–400. https://doi.org/10.1111/2041-210X.13115

McMurdie, P. J., & Holmes, S. (2013). phyloseq: An R Package for Reproducible Interactive Analysis and Graphics of Microbiome Census Data. PLOS ONE, 8(4), e61217. https://doi.org/10.1371/journal.pone.0061217

McMurdie, P. J., & Holmes, S. (2014). Waste Not, Want Not: Why Rarefying Microbiome Data Is Inadmissible. PLOS Computational Biology, 10(4), e1003531. https://doi.org/10.1371/journal.pcbi.1003531

146

Miliute, I., Buzaite, O., Baniulis, D., & Stanys, V. (2015). Bacterial endophytes in agricultural crops and their role in stress tolerance: A review. Zemdirbyste- Agriculture, 102, 465–478. https://doi.org/10.13080/z-a.2015.102.060

Mishra, M., Patole, S., & Mohapatra, H. (2017). Draft Genome Sequences of Nonclinical and Clinical Enterobacter cloacae Isolates Exhibiting Multiple Antibiotic Resistance and Virulence Factors. Genome Announcements, 5(45). https://doi.org/10.1128/genomeA.01218-17

Morella, N. M., Zhang, X., & Koskella, B. (2019). Tomato Seed-Associated Bacteria Confer Protection of Seedlings Against Foliar Disease Caused by Pseudomonas syringae. Phytobiomes Journal, 3(3), 177–190. https://doi.org/10.1094/PBIOMES-01-19-0007-R

Nandhini, S., Sendhilvel, V., & Babu, S. (2012). Endophytic bacteria from tomato and their efficacy against Fusarium oxysporum f. Sp. Lycopersici, the wilt pathogen. Journal of Biopesticides, 5(2), 178.

Nawangsih, A. A., Damayanti, I., Wiyono, S., & Kartika, J. G. (2011). Selection and Characterization of Endophytic Bacteria as Biocontrol Agents of Tomato Bacterial Wilt Disease. HAYATI Journal of Biosciences, 18(2), 66–70. https://doi.org/10.4308/hjb.18.2.66

Oksanen, J., Blanchet, F. G., Friendly, M., Kindt, R., Legendre, P., McGlinn, D., Minchin, P. R., O’Hara, R. B., Simpson, G. L., Solymos, P., Stevens, M. H. H., Szoecs, E., & Wagner, H. (2019). vegan: Community Ecology Package (Version R package version 2.5-6) [Computer software]. https://CRAN.R- project.org/package=vegan

Omeira, N., Barbour, E. K., Nehme, P. A., Hamadeh, S. K., Zurayk, R., & Bashour, I. (2006). Microbiological and chemical properties of litter from different chicken types and production systems. Science of The Total Environment, 367(1), 156– 162. https://doi.org/10.1016/j.scitotenv.2006.02.019

Ottesen, A. R., González Peña, A., White, J. R., Pettengill, J. B., Li, C., Allard, S., Rideout, S., Allard, M., Hill, T., Evans, P., Strain, E., Musser, S., Knight, R., & Brown, E. (2013). Baseline survey of the anatomical microbial ecology of an important food plant: Solanum lycopersicum (tomato). BMC Microbiology, 13(1), 114. https://doi.org/10.1186/1471-2180-13-114

Pan, X., Richardson, M. D., Deng, S., Kremer, R. J., English, J. T., Mihail, J. D., Sams, C. E., Scharf, P. C., Veum, K. S., & Xiong, X. (2017). Effect of Organic Amendment and Cultural Practice on Large Patch Occurrence and Soil

147

Microbial Community. Crop Science, 57(4), 2263–2272. https://doi.org/10.2135/cropsci2016.09.0809

Panke-Buisse, K., Poole, A. C., Goodrich, J. K., Ley, R. E., & Kao-Kniffin, J. (2015). Selection on soil microbiomes reveals reproducible impacts on plant function. The ISME Journal, 9(4), 980–989. https://doi.org/10.1038/ismej.2014.196

Paulson, J. N., Stine, O. C., Bravo, H. C., & Pop, M. (2013). Differential abundance analysis for microbial marker-gene surveys. Nature Methods, 10(12), 1200– 1202. https://doi.org/10.1038/nmeth.2658

Pavoine, S., Dufour, A.-B., & Chessel, D. (2004). From dissimilarities among species to dissimilarities among communities: A double principal coordinate analysis. Journal of Theoretical Biology, 228(4), 523–537. https://doi.org/10.1016/j.jtbi.2004.02.014

Peralta, A. L., Sun, Y., McDaniel, M. D., & Lennon, J. T. (2018). Crop rotational diversity increases disease suppressive capacity of soil microbiomes. Ecosphere, 9(5), e02235. https://doi.org/10.1002/ecs2.2235

Postma-Blaauw, M. B., Goede, R. G. M. de, Bloem, J., Faber, J. H., & Brussaard, L. (2010). Soil biota community structure and abundance under agricultural intensification and extensification. Ecology, 91(2), 460–473. https://doi.org/10.1890/09-0666.1

Quast, C., Pruesse, E., Yilmaz, P., Gerken, J., Schweer, T., Yarza, P., Peplies, J., & Glöckner, F. O. (2013). The SILVA ribosomal RNA gene database project: Improved data processing and web-based tools. Nucleic Acids Research, 41(D1), D590–D596. https://doi.org/10.1093/nar/gks1219

R Core Team. (2014). R: A language and environment for statistical computing. R Foundation for Statistical Computing. http://www.R-project.org/

Riegel, C., & Noe, J. P. (2000). Chicken Litter Soil Amendment Effects on Soilborne Microbes and Meloidogyne incognita on Cotton. Plant Disease, 84(12), 1275– 1281. https://doi.org/10.1094/PDIS.2000.84.12.1275

Romero, F. M., Marina, M., & Pieckenstain, F. L. (2014). The communities of tomato (Solanum lycopersicum L.) leaf endophytic bacteria, analyzed by 16S-ribosomal RNA gene pyrosequencing. FEMS Microbiology Letters, 351(2), 187–194. https://doi.org/10.1111/1574-6968.12377

Rosenzweig, N., Tiedje, J. M., Quensen, J. F., Meng, Q., & Hao, J. J. (2011). Microbial Communities Associated with Potato Common Scab-Suppressive Soil 148

Determined by Pyrosequencing Analyses. Plant Disease, 96(5), 718–725. https://doi.org/10.1094/PDIS-07-11-0571

Russelle, M. P., Entz, M. H., & Franzluebbers, A. J. (2007). Reconsidering integrated crop–livestock systems in North America. Agronomy Journal, 99(2), 325–334.

Sahoo, U. K., Singh, S. L., Gogoi, A., Kenye, A., & Sahoo, S. S. (2019). Active and passive soil organic carbon pools as affected by different land use types in Mizoram, Northeast India. PLoS ONE, 14(7). https://doi.org/10.1371/journal.pone.0219969

Saleem, M., Law, A. D., Sahib, M. R., Pervaiz, Z. H., & Zhang, Q. (2018). Impact of root system architecture on rhizosphere and root microbiome. Rhizosphere, 6, 47–51. https://doi.org/10.1016/j.rhisph.2018.02.003

Sanders, E. R., & Miller, J. H. (2010). I, Microbiologist: A Discovery-Based Course In Microbial Ecology and Molecular Evolution. ASM Press.

Santoyo, G., Moreno-Hagelsieb, G., del Carmen Orozco-Mosqueda, Ma., & Glick, B. R. (2016). Plant growth-promoting bacterial endophytes. Microbiological Research, 183, 92–99. https://doi.org/10.1016/j.micres.2015.11.008

Schiere, H., & Kater, L. (2001). Mixed crop-livestock farming. A review of traditional technologies based on literature and field experience. FAO Animal Production and Health Paper (FAO). https://agris.fao.org/agris- search/search.do?recordID=XF2003410307

Schliep, K., Potts, A. J., Morrison, D. A., & Grimm, G. W. (2017). Intertwining phylogenetic trees and networks. Methods in Ecology and Evolution, 8(10), 1212–1220. https://doi.org/10.1111/2041-210X.12760

Schneider, C. A., Rasband, W. S., & Eliceiri, K. W. (2012). NIH Image to ImageJ: 25 years of image analysis. Nature Methods, 9(7), 671–675. https://doi.org/10.1038/nmeth.2089

Shahzad, R., Khan, A. L., Bilal, S., Asaf, S., & Lee, I.-J. (2017). Plant growth- promoting endophytic bacteria versus pathogenic infections: An example of Bacillus amyloliquefaciens RWL-1 and Fusarium oxysporum f. sp. lycopersici in tomato. PeerJ, 5, e3107. https://doi.org/10.7717/peerj.3107

Shaik, S. P., & Thomas, P. (2019). In Vitro Activation of Seed-Transmitted Cultivation- Recalcitrant Endophytic Bacteria in Tomato and Host–Endophyte Mutualism. Microorganisms, 7(5). https://doi.org/10.3390/microorganisms7050132

149

Shannon, C. E. (1948). A Mathematical Theory of Communication. Bell System Technical Journal, 27(4), 623–656. https://doi.org/10.1002/j.1538- 7305.1948.tb00917.x

Sharon, J. A., Hathwaik, L. T., Glenn, G. M., Imam, S. H., & Lee, C. C. (2016). Isolation of efficient phosphate solubilizing bacteria capable of enhancing tomato plant growth. Journal of Soil Science and Plant Nutrition, 16(2), 525– 536. https://doi.org/10.4067/S0718-95162016005000043

Simon, H. M., Smith, K. P., Dodsworth, J. A., Guenthner, B., Handelsman, J., & Goodman, R. M. (2001). Influence of Tomato Genotype on Growth of Inoculated and Indigenous Bacteria in the Spermosphere. Applied and Environmental Microbiology, 67(2), 514–520. https://doi.org/10.1128/AEM.67.2.514-520.2001

Sogin, M. L., Morrison, H. G., Huber, J. A., Mark Welch, D., Huse, S. M., Neal, P. R., Arrieta, J. M., & Herndl, G. J. (2006). Microbial diversity in the deep sea and the underexplored “rare biosphere.” Proceedings of the National Academy of Sciences of the United States of America, 103(32), 12115–12120. https://doi.org/10.1073/pnas.0605127103

Soil Fertility Lab, O. S. U. (2019). Procedures for the Determination of Permanganate Oxidizable Carbon, Autoclaved-Citrate Extractable (ACE) Soil Protein and Soil Respiration. https://soilfertility.osu.edu/protocols

Soil Survey Staff. (2019). Natural Resources Conservation Service, United States Department of Agriculture. http://websoilsurvey.sc.egov.usda.gov/

Souza, S. A., Xavier, A. A., Costa, M. R., Cardoso, A. M. S., Pereira, M. C. T., & Nietsche, S. (2013). Endophytic bacterial diversity in banana ‘Prata Anã’ (Musa spp.) roots. Genetics and Molecular Biology, 36(2), 252–264. https://doi.org/10.1590/S1415-47572013000200016

Sulc, R. M., & Franzluebbers, A. J. (2014). Exploring integrated crop–livestock systems in different ecoregions of the United States. European Journal of Agronomy, 57, 21–30. https://doi.org/10.1016/j.eja.2013.10.007

Sun, H. Y., Deng, S. P., & Raun, W. R. (2004). Bacterial Community Structure and Diversity in a Century-Old Manure-Treated Agroecosystem. Applied and Environmental Microbiology, 70(10), 5868–5874. https://doi.org/10.1128/AEM.70.10.5868-5874.2004

Tamošiūnė, I., Stanienė, G., Haimi, P., Stanys, V., Rugienius, R., & Baniulis, D. (2018). Endophytic Bacillus and Pseudomonas spp. Modulate Apple Shoot Growth, 150

Cellular Redox Balance, and Protein Expression Under in Vitro Conditions. Frontiers in Plant Science, 9. https://doi.org/10.3389/fpls.2018.00889

Tamura, K., Nei, M., & Kumar, S. (2004). Prospects for inferring very large phylogenies by using the neighbor-joining method. Proceedings of the National Academy of Sciences of the United States of America, 101(30), 11030–11035. https://doi.org/10.1073/pnas.0404206101

Telias, A., White, J. R., Pahl, D. M., Ottesen, A. R., & Walsh, C. S. (2011). Bacterial community diversity and variation in spray water sources and the tomato fruit surface. BMC Microbiology, 11, 81. https://doi.org/10.1186/1471-2180-11-81

Thokchom, E., Thakuria, D., Kalita, M. C., Sharma, C. K., & Talukdar, N. C. (2017). Root colonization by host-specific rhizobacteria alters indigenous root endophyte and rhizosphere soil bacterial communities and promotes the growth of mandarin orange. European Journal of Soil Biology, 79, 48–56. https://doi.org/10.1016/j.ejsobi.2017.02.003

Tian, B., Zhang, C., Ye, Y., Wen, J., Wu, Y., Wang, H., Li, H., Cai, S., Cai, W., Cheng, Z., Lei, S., Ma, R., Lu, C., Cao, Y., Xu, X., & Zhang, K. (2017). Beneficial traits of bacterial endophytes belonging to the core communities of the tomato root microbiome. Agriculture Ecosystems & Environment, 247, 149–156. https://doi.org/10.1016/j.agee.2017.06.041

Toju, H., Okayasu, K., & Notaguchi, M. (2019). Leaf-associated microbiomes of grafted tomato plants. Scientific Reports, 9(1), 1787. https://doi.org/10.1038/s41598-018-38344-2

Udo, H. M. J., Aklilu, H. A., Phong, L. T., Bosma, R. H., Budisatria, I. G. S., Patil, B. R., Samdup, T., & Bebe, B. O. (2011). Impact of intensification of different types of livestock production in smallholder crop-livestock systems. Livestock Science, 139(1), 22–29. https://doi.org/10.1016/j.livsci.2011.03.020

USDA-AMS. (2019). USDA: Agricultural Marketing Service. Agricultural Marketing Service. https://www.ams.usda.gov/

Wagner, M. R., Lundberg, D. S., Del Rio, T. G., Tringe, S. G., Dangl, J. L., & Mitchell- Olds, T. (2016). Host genotype and age shape the leaf and root microbiomes of a wild perennial plant. Nature Communications, 7, 12151. https://doi.org/10.1038/ncomms12151

Walker, J. J., & Pace, N. R. (2007). Phylogenetic Composition of Rocky Mountain Endolithic Microbial Ecosystems. Applied and Environmental Microbiology, 73(11), 3497–3504. https://doi.org/10.1128/AEM.02656-06 151

Walpola, B., & Yoon, M.-H. (2013). Isolation and characterization of phosphate solubilizing bacteria and their co-inoculation efficiency on tomato plant growth and phosphorous uptake. Afr J Microbiol Res, 7, 266–275. https://doi.org/10.5897/AJMR12.2282

Walterson, A. M., & Stavrinides, J. (2015). Pantoea: Insights into a highly versatile and diverse genus within the Enterobacteriaceae. FEMS Microbiology Reviews, 39(6), 968–984. https://doi.org/10.1093/femsre/fuv027

Wander, M. (2004). Soil Organic Matter Fractions and Their Relevance to Soil Function. In F. Magdoff & R. Weil (Eds.), Soil Organic Matter in Sustainable Agriculture (Vol. 20042043). CRC Press. https://doi.org/10.1201/9780203496374.ch3

Wander, M. M., Traina, S. J., Stinner, B. R., & Peters, S. E. (1994). Organic and Conventional Management Effects on Biologically Active Soil Organic Matter Pools. Soil Science Society of America Journal, 58(4), 1130–1139. https://doi.org/10.2136/sssaj1994.03615995005800040018x

Wang, Q., Garrity, G. M., Tiedje, J. M., & Cole, J. R. (2007). Naïve Bayesian Classifier for Rapid Assignment of rRNA Sequences into the New Bacterial Taxonomy. Applied and Environmental Microbiology, 73(16), 5261–5267. https://doi.org/10.1128/AEM.00062-07

Weisburg, W. G., Barns, S. M., Pelletier, D. A., & Lane, D. J. (1991). 16S ribosomal DNA amplification for phylogenetic study. Journal of Bacteriology, 173(2), 697–703. https://doi.org/10.1128/jb.173.2.697-703.1991

Wickham, H., Hester, J., & Francois, R. (2018). readr: Read Rectangular Text Data. (R package version 1.3.1.) [Computer software]. https://CRAN.R- project.org/package=readr

Willis, A., & Bunge, J. (2015). Estimating diversity via frequency ratios. Biometrics, 71(4), 1042–1049. https://doi.org/10.1111/biom.12332

Woods, L. E. (1989). Active organic matter distribution in the surface 15 cm of undisturbed and cultivated soil. Biology and Fertility of Soils, 8(3). https://doi.org/10.1007/BF00266490

Wright, E., S. (2016). Using DECIPHER v2.0 to Analyze Big Biological Sequence Data in R. The R Journal, 8(1), 352. https://doi.org/10.32614/RJ-2016-025

152

Xia, Y., DeBolt, S., Dreyer, J., Scott, D., & Williams, M. A. (2015). Characterization of culturable bacterial endophytes and their capacity to promote plant growth from plants grown using organic or conventional practices. Frontiers in Plant Science, 6. https://doi.org/10.3389/fpls.2015.00490

Ye, J., Zhang, R., Nielsen, S., Joseph, S. D., Huang, D., & Thomas, T. (2016). A Combination of Biochar–Mineral Complexes and Compost Improves Soil Bacterial Processes, Soil Quality, and Plant Properties. Frontiers in Microbiology, 7. https://doi.org/10.3389/fmicb.2016.00372

Zazou, K., Abdelkrim, T., Chama, Z., Kaldi, A., Tounssi, S., & Bouziane, A. (2016). Isolation and characterization of Microbacterium arthrosphaerae and its effect on tomato culture growth. Der Pharm. Lett., 8(8), 1–8.

Zhang, D., Yan, D., Cheng, H., Fang, W., Huang, B., Wang, X., Wang, X., Yan, Y., Ouyang, C., Li, Y., Wang, Q., & Cao, A. (2020). Effects of multi-year biofumigation on soil bacterial and fungal communities and strawberry yield. Environmental Pollution, 256, 113415. https://doi.org/10.1016/j.envpol.2019.113415

Zhang, L., Li, L., Pan, X., Shi, Z., Feng, X., Gong, B., Li, J., & Wang, L. (2018). Enhanced Growth and Activities of the Dominant Functional Microbiota of Chicken Manure Composts in the Presence of Maize Straw. Frontiers in Microbiology, 9. https://doi.org/10.3389/fmicb.2018.01131

Zhang, M., Powell, C. A., Guo, Y., Benyon, L., & Duan, Y. (2013). Characterization of the microbial community structure in Candidatus Liberibacter asiaticus-infected citrus plants treated with antibiotics in the field. BMC Microbiology, 13(1), 112. https://doi.org/10.1186/1471-2180-13-112

153

Tables

Soil health measurements1

POXC2 Soil protein Respiration Strategy P P P Mean Mean Mean value value value Chicken Grazed 765.62 8.28a 279.37a Pasture 0.56 0.08 . 0.04 * Pasture Only 711.16 7.32b 210.67b

* significant at p = 0.05 . significant at p < 0.1 1Log 10 transformation was done on all measurements that did not meet parametric assumptions due to non-normality and non-homogeneity of variance. 2Permanganate oxidizable carbon.

Table 3.1. Effect of chicken grazing history on active pools of organic matter in soil collected from plots prior to planting of the tomato.

154

Sum of Mean F Source of Variation df Pr(>F) Squares Square value

Aboveground Biomass1 Rep 3 0.03 0.01 0.07 0.97 Block 3 2.93 0.98 7.96 0.01 ** Error (a) (Main Plot) 9 1.11 0.12 Treatment 1 0.02 0.02 0.29 0.60 Block:Treatment 3 0.24 0.08 1.38 0.30 Error (b) (Subplot) 12 0.68 0.06

Number of Compound Leaves Rep 3 2.63 0.88 0.40 0.76 Block 3 58.38 19.46 8.81 0.00 ** Error (a) (Main Plot) 9 19.88 2.21 Treatment 1 0.13 0.13 0.06 0.80 Block:Treatment 3 10.38 3.46 1.77 0.21 Error (b) (Subplot) 12 23.50 1.96

*** significant at P=0 ** significant at P=0.001 1Data was transformed due to non-normality.

Table 3.2. Impact of block and chicken-grazing history on aboveground biomass and number of compound leaves of tomato plants sampled in July.

155

Aboveground Number of Compound

Biomass Leaves Treatment Standard Standard Mean Mean Deviation Deviation Chicken Grazed 13.77 Pasture 19.10 7.25 1.77 Pasture Only 19.42 15.38 7.38 2.12

Block1

105 8.49 5.88a 6.38 1.60a 201 29.39 12.87b 8.38 1.06b 301 30.90 12.17b 8.88 1.25b 403 8.26 4.74a 5.63 1.69a

1Each block represents a whole plot.

Table 3.3. Impact of block and chicken-grazing history on aboveground biomass and number of compound leaves of tomato plants sampled in July. Means in a column without a common superscript letter differ (p < 0.05), as analyzed by Fisher’s LSD test.

156

Sum of Mean Source of Variation df F value Pr(>F) Squares Square

Aboveground Biomass1 Rep 3 0.10 0.03 0.66 0.60 Block 3 0.91 0.30 5.77 0.02 * Error (a) (Main Plot) 9 0.47 0.05 Treatment 1 0.06 0.06 1.54 0.24 Block:Treatment 3 0.17 0.06 1.38 0.30 Error (b) (Subplot) 12 0.49 0.04

Number of Compound Leaves1 Rep 3 0.02 0.01 0.18 0.91 Block 3 0.25 0.08 2.77 0.10 Error (a) (Main Plot) 9 0.27 0.03 Treatment 1 0.03 0.03 1.92 0.19 Block:Treatment 3 0.06 0.02 1.33 0.31 Error (b) (Subplot) 12 0.17 0.01

Total Biomass of Tomatos Harvested1 Rep 3 0.29 0.10 1.24 0.35 Block 3 0.69 0.23 3.00 0.09 . Error (a) (Main Plot) 9 0.70 0.08 Treatment 1 0.01 0.01 0.04 0.85 Block:Treatment 3 0.07 0.02 0.14 0.94 Error (b) (Subplot) 12 2.01 0.17

Continued *** significant at P=0 ** significant at P=0.001 . significant at P < 0.1 1Data was transformed due to non-normality.

Table 3.4. Impact of block and chicken-grazing history on above-ground biomass, number of compound leaves, total biomass of tomatoes harvested, number of symptomatic tomatoes harvested, number of total tomatoes harvested and root biomass of tomato plants sampled in September.

157

Table 3.4 Continued

Sum of Mean Source of Variation df F value Pr(>F) Squares Square

Number of Symptomatic Tomatoes Harvested1 Rep 3 0.16 0.05 3.50 0.06 . Block 3 0.71 0.24 15.53 6.65x10-4 *** Error (a) (Main Plot) 9 0.14 0.02 Treatment 1 0.01 0.01 0.03 0.86 Block:Treatment 3 0.04 0.01 0.08 0.97 Error (b) (Subplot) 12 2.02 0.17

Number of Total Tomatoes Harvested1 Rep 3 0.21 0.07 2.12 0.17 Block 3 0.60 0.20 6.01 0.02 * Error (a) (Main Plot) 9 0.30 0.03 Treatment 1 0.02 0.02 0.16 0.69 Block:Treatment 3 0.07 0.02 0.20 0.90 Error (b) (Subplot) 12 1.41 0.12

Root Biomass1 Rep 3 0.01 2.10x10-3 0.08 0.97 Block 3 1.22 0.41 13.97 9.81x10-4 *** Error (a) (Main Plot) 9 0.26 0.03 Treatment 1 0.06 0.06 2.61 0.13 Block:Treatment 3 0.04 0.01 0.54 0.67 Error (b) (Subplot) 12 0.27 0.02

*** significant at P=0 ** significant at P=0.001 . significant at P < 0.1 1Data was transformed due to non-normality

158

Treatment Block1 Measurements Chicken Pasture Only 105 201 301 403 Grazed Pasture Mean 128.48 89.18 68.42 175.42 121.15 70.33 Aboveground Biomass Standard Deviation 98.05 39 34.92a 96.24b 72.98ab 31.90a

Number of Compound Mean 34 28 25.63 40.38 33.75 24.25 Leaves Standard Deviation 15.36 7.9 8.80a 11.11b 15.89ab 5.75a

Total Biomass of Mean 617.42 540.35 297.50 826.71 708.92 482.43 Tomatoes Harvested Standard Deviation 526.07 368.58 128.46 475.85 655.67 244.13

Number of Mean 4.00 3.81 2.25 5.38 5.63 2.38 159 Symptomatic Standard Deviation 3.36 3.06 1.04 3.5 4.27 1.06 Tomatoes

Number of Total Mean 6.19 5.44 3.25 8.00 7.50 4.50 Tomatoes Harvested Standard Deviation 4.89 3.67 1.28a 4.87b 5.78ab 1.93a

Mean 16.72 13.97 9.74 22.21 21.04 8.37 Root Biomass Standard Deviation 9.27 7.42 4.51a 4.73b 8.93b 2.47a

1Each block represents a whole plot.

Table 3.5. Impact of block and chicken-grazing history on aboveground biomass, number of compound leaves, total biomass of tomatoes harvested, number of symptomatic tomatoes, number of total tomatoes harvested and root biomass of tomato plants sampled in September. Means in a row without a common superscript letter differ (p < 0.1), as analyzed by Fisher’s LSD test.

159

Sampling Plant Number of Month Tissue Treatment Isolates Closest Match in NCBI Identity Recovered Isolate ID1 Recovered Recovered with Exact Database (Accession Number) %3 J SP S TF CG PO Sequences2

Bacillus 1 (AVC.025) 9 31 26 14 20 20 40 Bacillus sp. (MN969919.1) 99.71% Bacillus 2 (AVC.038) 4 4 4 4 Bacillus sp. (MH683121.1) 99.71% Bacillus 3 (AVC.051) 5 6 8 3 6 5 11 Bacillus sp. (MN865961.1) 99.93% Bacillus 4 (AVC.054) 2 4 2 4 2 4 6 Bacillus sp. (MN966966.1) 99.85% Bacillus 10 (AVC.146) 1 1 1 1 Bacillus sp. (MN793064.1) 100% Chryseobacterium 1 Chryseobacterium sp. 1 1 1 1 97.57% (AVC.135) (MN081010.1)

160 Comamonas 1 1 1 1 1 Comamonas sp. (KX461916.1) 100%

(AVC.114) Curtobacterium 1 Curtobacterium sp. 1 1 1 1 99.93% (AVC.177) (MK883133.1) Enterobacter 1 2 2 2 2 Enterobacter sp. (KP792627.1 ) 99.85% (AVC.156) Enterobacteriaceae 1 2 2 2 2 Enterobacter sp. (MN641475.1) 99.70% (AVC.057)

Continued 1AVC: Reference isolates from tomato endophyte culture collection. 2Recovered isolate sequence considered exactly the same were determined by computing pairwise distance (0.000; sequences were the same). 3The identity percent describes how similar the query sequence is to the target sequence. The higher the percent identity is, the more significant the match.

Table 3.6. Identification and distribution of bacterial endophytes recovered from the stem and tomato flesh of tomato plants grown in plots with and without history of chicken grazing. Abbreviations: J: July, SP: September, S: Stem, TF: Tomato flesh, CG: Chicken grazed, PO: Pasture only.

160

Table 3.6. Continued

Sampling Plant Number of Month Tissue Treatment Isolates Closest Match in NCBI Identity Recovered Isolate ID1 Recovered Recovered with Exact Database (Accession Number) %3 J SP S TF CG PO Sequences2

Frigoribacterium 1 Frigoribacterium sp. (AVC.125) 1 1 1 1 (MN989058.1) 99.93% Massilia 1 (AVC.234) 1 1 1 1 Massilia sp. (AM237367.1) 97.64% Microbacterium 1 Microbacterium sp. (AVC.022) 2 3 4 1 4 1 5 (JQ660113.1) 99.55% Microbacterium 5 Microbacterium sp. (AVC.066) 1 1 1 1 (MG016448.1) 100% Paenibacillus 3 161 (AVC.149) 1 1 1 1 Paenibacillus sp. (GU647099.1) 99.52% Paenibacillus 4 (AVC.172) 1 1 1 1 Paenibacillus sp. (KP279976.1) 99.63% Pantoea 1 (AVC.130) 1 1 1 1 1 1 2 Pantoea sp. (MK618629.1) 99.93% Pantoea 2 (AVC.218) 1 1 1 1 Pantoea sp. (CP048033.1) 99.11% Pseudomonas 1 Pseudomonas sp. (AVC.041) 8 21 24 5 17 12 29 (MN326729.1) 99.76% Pseudomonas 2 Pseudomonas sp. (AVC.116) 1 1 1 1 (MH769577.1) 99.21%

Continued 1AVC: Reference isolates from tomato endophyte culture collection. 2Recovered isolate sequence considered exactly the same were determined by computing pairwise distance (0.000; sequences were the same). 3The identity percent describes how similar the query sequence is to the target sequence. The higher the percent identity is, the more significant the match.

161

Table 3.6. Continued

Sampling Plant Number of Month Tissue Treatment Isolates Closest Match in NCBI Identity Recovered Isolate ID1 Recovered Recovered with Exact Database (Accession Number) %3 J SP S TF CG PO Sequences2

Pseudomonas 3 (AVC.154) 1 1 1 1 Pseudomonas sp. (AB720144.1) 99.86% Pseudomonas 4 (AVC.160) 1 1 1 1 Pseudomonas sp. (KU305727.1) 100% Pseudomonas 5 (AVC.169) 1 1 1 1 Pseudomonas sp. (KU305727.1) 100% Rhizobium 1 (AVC.221) 1 1 1 1 Rhizobium sp. (MG836173.1) 100% 162 Sphingobacterium 1 Sphingobacterium sp. (AVC.136) 1 1 1 1 (KY767658.1) 98.89% Sphingomonas 1 Sphingomonas sp. (AVC.153) 1 1 1 1 (MN932317.1) 98.73% Sphingomonas 2 Sphingomonas sp. (NR (AVC.157) 1 1 1 1 109167.1) 97.56% Sphingomonas 3 Sphingomonas sp. (AVC.203) 3 2 1 3 3 (MH769434.1) 100% Stenotrophomonas 1 Stenotrophomonas sp. (AVC.111) 1 1 2 2 2 (JN638427.1) 99.93%

1AVC: Reference isolates from tomato endophyte culture collection. 2Recovered isolate sequence considered exactly the same were determined by computing pairwise distance (0.000; sequences were the same). 3The identity percent describes how similar the query sequence is to the target sequence. The higher the percent identity is, the more significant the match.

162

Number After Origin of Total Reads / Forward/Reverse of Primer/Adapter Sample Reads1 Sample Sequence Samples Removal2 Reads 25~343,954 13~167,531 All samples 8,658,317 154 Average 56,222 27,458 Chicken Reads 21,649~127,185 10,580~62,473 271,411 4 Manure Average 68,103 33,475 Reads 4,458~21,210 2,201~10,384 Bulk Soil 98,475 7 Average 14,067 6,921 Rhizosphere Reads 3,763~101,284 1,808~49,847 1,332,727 36 Soil Average 37,020 18,189 Reads 2,191~338,413 1,039~164,552 Root 2,343,792 35 Average 66,965 32,738 Reads 35,669~343,954 17,360~167,531 Stem 2,831,813 32 Average 88,494 43,062 Reads 6,958~164555 3,438~80,445 Leaf 1,383,756 27 Average 51,250 25,020 Tomato Reads 27,525~79,945 13,386~38,633 229,887 4 Flesh Average 57,472 27,895 Tomato Reads 23,839~68,399 11,515~33,257 164,271 4 Seed Average 41,068 19,977 Reads 26~876 13~434 Controls 1,263 5 Average 253 112

1The total reads of all samples in 2019 before processing and filtering. 2Primer and adapter removal performed in Cutadapt software.

Table 3.7. Recovery of amplicon sequencing reads from different tomato plant tissues, rhizosphere soil and baseline samples (bulk soil and chicken manure). Using Cutadapt software, each primer and adapter was removed.

163

Origin of Reads / Quality filtered1 Denoised F2 Denoised R3 Merged Pairs4 Nonchimeric5 Sample Sample Reads 9~143,262 9~143,142 9~143,232 6~54,709 6~53,964 All samples Average 23,584 23,451 23,508 10,197 9,892 Chicken Reads 9,334~55,432 9,173~55,283 9,236~55,394 8,799~54,709 8,745~53,964 manure Average 29,593 29,497 29,544 29,095 28,781 Reads 1,927~8,919 1,923~8,737 1,924~8,815 1,901~7,247 1,900~7,197 Bulk soil Average 5,970 5,858 5,904 4,256 4,132 Rhizosphere Reads 1,559~42,574 1,513~41,777 1531~42,131 1,361~30,447 1,353~30,088 soil Average 15,593 15,245 15,384 11,489 11,122 Reads 898~139,197 896~138,860 894~139,025 847~53,785 845~49,826 164 Root Average 28,192 28,045 28,106 14,631 14,031 Reads 14,903~143,262 14,898~143,142 14,896~143,232 1,899~30,859 1,894~29,693 Stem Average 36,941 36,900 36,926 5,974 5,713 Reads 1,711~68,455 1,711~68,425 1,711~68,451 1,696~32,719 1,696~32,451 Leaf Average 21,352 21,343 21,351 10,313 10,251 Reads 11,674~32,553 11,656~32,489 11,667~32,526 1,037~8,841 1,033~8,590 Tomato flesh Average 24,070 24,041 24,060 3,778 3,687 Reads 9,306~28,615 9,302~28,606 9,305~28,615 291~8,971 285~8,821 Tomato seed Average 17,097 17,079 17,092 3,347 3,276 Reads 9~375 9~372 9~375 6~366 6~366 Controls Average 93 92 93 79 79

1-5Processing done using DADA2 software. Low-quality reads were filtered out, forward and reversereads were denoised and merged, chimeric reads were removed.

Table 3.8. Recovery of amplicon sequencing reads from different tomato plant tissues, rhizosphere soil and baseline samples (bulk soil and chicken manure). Using DADA2, reads were processed and filtered, per sample sequenced.

164

Before After Number of Origin of Reads / Removing Removing Samples After Sample Sample Unwanted Unwanted Removing Ones Taxa1 Taxa2 with 0 Reads Reads 0~53,988 0~53,759 148 All samples Average 7,108 6,372 Reads 8,754~53,988 8,641~53,749 4 Chicken Manure Average 28,838 28,730 Reads 685~7,189 167~7,107 7 Bulk Soil Average 3,913 3,788 Reads 1,375~30,464 1,360~30,081 35 Rhizosphere Soil Average 11,163 11,008 Reads 30~48,938 29~40,813 33 Root Average 12,941 9,967 Reads 15~8,848 15~8,467 32 Stem Average 2,240 2,184 Reads 0~1,841 0~1,790 24 Leaf Average 285 277 Reads 72~6,157 71~5,948 4 Tomato Flesh Average 2,158 2,101 Reads 17~8,732 17~8,732 4 Tomato Seed Average 2,226 2,224 Reads 6~142 6~48 5 Controls Average 35 16

1-2Unwanted taxa removed belong to chloroplast, mitochondria and eukaryotic DNA.

Table 3.9. Recovery of amplicon sequencing reads from different tomato plant tissues, rhizosphere soil and baseline samples (bulk soil and chicken manure). Using phyloseq, reads belonging to unwanted taxa were removed.

165

Indices means by treatment Sample origin and indices Chicken grazed Pasture only Bulk soil1 Chao1 Species Richness 4848.68 3277.01 Shannon's Diversity Index 7.29 5.37 Faith's Phylogenetic Diversity 116.98 83.61

Chicken Manure2 Chao1 Species Richness 423.99 Shannon's Diversity Index 3.00 Faith's Phylogenetic Diversity 23.24

1There were four samples for chicken grazed pasture plots and three samples from pasture only plots. Bulk soil samples were collected in May. 2There were four chicken manure samples, all collected in September.

Table 3.10. Increase in bacterial community richness, evenness and diversity in samples obtained from chicken-grazed pasture plots. Alpha diversity metrics (Chao1 species richness, Shannon’s diversity index and Faith’s Phylogenetic diversity) of bacterial communities from bulk soil and chicken manure collected in May and September, respectively.

166

Indices Means by Sampling Month and Treatment Sample Origin and July September Indices Chicken Sample Pasture Sample Chicken Sample Pasture Sample

Grazed Size Only Size Grazed Size Only Size Rhizosphere Soil Chao1 Species Richness 5404.69 5166.66 3952.09 5336.60 Shannon's Diversity Index 6.77 9 7.07 10 5.70 8 6.99 9 Faith's Phylogenetic Diversity 162.23 166.72 100.77 120.41

Root Chao1 Species Richness 2131.27 1955.61 750.40 1347.76 Shannon's Diversity Index 4.67 10 5.31 9 3.13 8 3.08 8 167 Faith's Phylogenetic Diversity 68.32 69.40 35.93 43.34

Stem Chao1 Species Richness 1700.66 1682.45 191.66 1614.77 Shannon's Diversity Index 4.76 8 4.46 8 3.20 7 4.37 7 Faith's Phylogenetic Diversity 17.36 13.23 53.79 88.65

Leaf Chao1 Species Richness 3217.36 510.84 3531.28 1614.77 Shannon's Diversity Index 4.99 8 3.16 8 4.58 4 4.37 2 Faith's Phylogenetic Diversity 63.30 15.60 144.40 88.65

Table 3.11. Variations in bacterial community richness, evenness and diversity in samples obtained from chicken-grazed pasture plots and pasture only plots. Alpha diversity metrics (Chao1 species richness, Shannon’s diversity index and Faith’s Phylogenetic diversity) of bacterial communities from rhizosphere soil, root, stem and leaf of tomato plants sampled in June and Septem

167

Indices Means by Treatment Sample Origin and Indices1 Chicken Grazed Pasture Only Tomato Flesh Chao1 Species Richness 130.85 495.74 Shannon's Diversity Index 2.13 3.46 Faith's Phylogenetic Diversity 10.48 26.43

Tomato Seed Chao1 Species Richness 271.03 190.72 Shannon's Diversity Index 3.45 2.15 Faith's Phylogenetic Diversity 0.92 0.71

1There were four samples for both sample origins per treatment. Tomato flesh and seed was collected in September.

Table 3.12. Variations in bacterial community richness, evenness and diversity in tomato fruit and seed samples obtained from chicken-grazed pasture plots and pasture only plots. Alpha diversity metrics (Chao1 species richness, Shannon’s diversity index and Faith’s Phylogenetic diversity) of bacterial communities from tomato flesh and seed sampled in September.

168

Indices Means by Sampling Month and Treatment July Sample Origin and Chicken Grazed Pasture Only P Indices Value1 Sample Sample Sample Sample 201 301 201 301 Size size Size Size Rhizosphere soil Chao1 Species Richness 7450.49 3768.05 6547.41 3095.54 <0.001 Shannon's Diversity 7.73 6.01 7.60 6.28 <0.0004 Index 4 5 6 3 Faith's Phylogenetic 195.48 135.63 145.90 197.93 <0.002 Diversity 169

Leaf Chao1 Species Richness 95.26 6339.46 136.23 1178.61 <0.008 Shannon's Diversity 2.45 7.54 2.67 4.50 <0.003 Index 4 4 4 4 Faith's Phylogenetic 6.30 120.31 7.57 30.01 Diversity <0.007 Continued

1Kruskal Wallis test was used to determine significant differences between plots between sampling months. P value is considered significant at 0.1.

Table 3.13. Alpha diversity metrics (Chao1 species richness, Shannon’s diversity index and Faith’s Phylogenetic diversity) of bacterial communities from rhizosphere soil and leaf from both sampling months and treatments.

169

Table 3.13 Continued

Indices means by Sampling Month and Treatment September P Sample Origin and Indices Chicken Grazed Pasture Only Value1 Sample Sample Sample Sample 201 301 201 301 Size Size Size Size Rhizosphere soil Chao1 Species Richness 5869.80 2034.37 5945.03 4849.86 <0.001

170 Shannon's Diversity Index 7.61 4 3.79 4 3.79 4 6.54 5 <0.0004

Faith's Phylogenetic Diversity 91.07 110.47 110.47 162.68 <0.002

Leaf Chao1 Species Richness 29.88 29.88 25.25 <0.008 Shannon's Diversity Index 1.43 2 1.43 2 1.39 2 <0.003 Faith's Phylogenetic Diversity 139.62 139.62 7.74 <0.007

1Kruskal Wallis test was used to determine significant differences between plots between sampling months. P value is considered significant at 0.1.

170

Sample Number of Number of Origin Phylum and Corresponding Differentially ASVs (Sampling Families Abundant Considered Month) ASVs Rhizosphere Soil (July) 16,192 Acidobacteria 3 NA 2 Thermoanaerobaculaceae 1 Actinobacteria 16 67-14 4 1 Gaiellaceae 1 1 Microbacteriaceae 2 Mycobacteriaceae 2 NA 1 1 Solirubrobacteraceae 3 Bacteroidetes 3 Chitinophagaceae 2 Saprospiraceae 1 1 Roseiflexaceae 1 Firmicutes 8 Alicyclobacillaceae 7 Family_XII 1 1 Gemmatimonadaceae 1 2 Nitrospiraceae 2 Proteobacteria 18 A21b 2 Blfdi19 1 Continued Table 3.14. Differentially abundant taxa (p<0.1) between chicken grazed pasture plots and pasture only plots within sampling months. All taxa shown, except (indicate which) were more abundant in chicken grazed plots.

171

Table 3.14 Continued Sample Number of Number of Origin Phylum and Corresponding Differentially ASVs (Sampling Families Abundant Considered Month) ASVs Rhizosphere Soil (July) 16,192 Burkholderiaceae 1 Dongiaceae 1 Hyphomicrobiaceae 1 Micropepsaceae 1 Nitrosomonadaceae 1 Rhizobiaceae 1 Rhizobiales_Incertae_Sedis 1 Rhodomicrobiaceae 1 Sphingomonadaceae 2 TRA3-20 1 Unknown_Family 1 Xanthobacteraceae 2 Xanthomonadaceae 1 Verrucomicrobia 1 Pedosphaeraceae 1

Total 53

Stem (July) 1,615 Actinobacteria 2 Kineosporiaceae 2 Deinococcus-Thermus 1 Deinococcaceae 1 Proteobacteria 12 Acetobacteraceae 1 Beijerinckiaceae 1 Burkholderiaceae 6 Rhizobiaceae 1 Sphingomonadaceae 3

Total 15 Continued 172

Table 3.14 Continued

Sample Number of Number of Origin Phylum and Corresponding Differentially ASVs (Sampling Families Abundant Considered Month) ASVs Root (September) 9,397 Proteobacteria 1 Burkholderiaceae 1

Total 1

173

Figures

174

Figure 3.1. Overview of the field trial conducted in the Summer of 2019 at Mellinger Research Farm. Plots grazed by chickens in August 2017 and 2018 were used as vegetable plots in 2019. Tomato plants were grown in soils with differing chicken grazing history.

174

175

Figure 3.2. Overview of the three sampling months, involving the bulk soil, tomato plant and chicken manure sampling. At each sampling time, different samples were collected and analyzed for bacterial diversity, plant vigor and soil characteristics.

175

176

Figure 3.3. Soil sampling and homogenization for subsequent use in amplicon sequencing, as planting substrate and soil health tests. Soil obtained from chicken-grazed pasture plots and pasture only plots at Mellinger Research Farm. Soil was bagged in separate plastic bags per split plot. In the greenhouse, soil was poured into plastic containers, separated by treatment and plot, and mixed- together and homogenized by hand.

176

177

Figure 3.4. Three-week old tomato seedling transplant. On June 14th, 10 tomato seedlings were selected per treatment and plot to be transplanted to the corresponding split plot were soil was originally collected from.

177

Figure 3.5. Amplicon sequence data processing and analysis pipeline adapted from Huo, 2020.

178

18 Bacillus 1 (AVC.025) 16 Bacillus 10 (AVC.146) Bacillus 2 (AVC.038) 14 Bacillus 3 (AVC.051) Bacillus 4 (AVC.054) Chryseobacterium 1 (AVC.135) 12 Comamonas 1 (AVC.114) Curtobacterium 1 (AVC.177) 10 Enterobacteriaceae 1 (AVC.057) Enterobacter 1 (AVC.156) 8 Massilia 1 (AVC.234) Microbacterium 1 (AVC.022) Recovered Microbacterium 5 (AVC.066) 6 Paenibacillus 4 (AVC.172) Pantoea 2 (AVC.218) 4 Pseudomonas 1 (AVC.041) Pseudomonas 2 (AVC.116) Number of Bacterial Number of Endophytes 2 Pseudomonas 3 (AVC.154) Pseudomonas 4 (AVC.160) 0 Pseudomonas 5 (AVC.169) Chicken Grazed Pasture Only Chicken Grazed Pasture Only Chicken Grazed Pasture Only Chicken Grazed Pasture Only Rhizobium 1 (AVC.221) Pasture Pasture Pasture Pasture Sphingobacterium 1 (AVC.136) Plot 201 . Plot 301 . Plot 403 . Plot 105 Sphingomonas 1 (AVC.153) 179 Plots and Treatments

Figure 3.6. Endophytic bacterial isolates recovered from stem tissue of tomato plants grown in plots with or without history of chicken grazing. Different colors indicate the taxa to which each isolate belongs.

179

10 9 8 7 Bacillus 1 (AVC.025) 6 Bacillus 2 (AVC.038) Bacillus 3 (AVC.051) 5 Bacillus 4 (AVC.054) 4 Bacillus 10 (AVC.146)

Recovered Microbacterium 1 (AVC.022) 3 Pantoea 1 (AVC.130) 2 Pseudomonas 1 (AVC.041) Sphingomonas 2 (AVC.157) 1 Sphingomonas 3 (AVC.203)

Number of Bacterial Number of Endophytes 0 Chicken Grazed Pasture Only Chicken Grazed Pasture Only Chicken Grazed Pasture Only Chicken Grazed Pasture Only Pasture Pasture Pasture Pasture Plot 201 . Plot 301 . Plot 403 . Plot 105 Plos and Treatments

Figure 3.7. Endophytic bacterial isolates recovered from tomato flesh of tomatoes harvested from plants grown in plots with or

180 without history of chicken grazing. Different colors indicate the taxa to which each isolate belongs.

180

Figure 3.8. Sequencing depth for each sample origin displayed using rarefaction curves. Each sample is divided by treatment. Legend represents sampling months.

181

182

Figure 3.9. The relative abundance of major phyla in bulk soil samples collected in May. Each bar represents a sample within a treatment.

182

183

Figure 3.10. The relative abundance of major phyla in tomato root samples. Each bar represents a sample within a treatment in a sampling month.

183

184

Figure 3.11. The relative abundance of major phyla in tomato stem samples. Each bar represents a sample within a treatment in a sampling month.

184

185

Figure 3.12. The relative abundance of major phyla in tomato leaf samples. Each bar represents a sample within a treatment in a sampling month.

185

Figure 3.13. The relative abundance of major phyla in tomato flesh samples. Each bar represents a sample within a treatment in a sampling month.

186

Figure 3.14. The relative abundance of major phyla in tomato seed samples. Each bar represents a sample within a treatment in a sampling month.

187

188

Figure 3.15. The relative abundance of major phyla in tomato rhizosphere soil samples. Each bar represents a sample within a treatment in a sampling month.

188

Figure 3.16. The relative abundance of major phyla in chicken manure samples. Each bar represents a sample within a plot.

189

Figure 3.17. Bacterial communities were significantly separated by sampling month and sample origin. Principal coordinate analysis of bacterial communities of rhizosphere soil, root and stem samples. Weighted normalized UniFrac distance matrix was calculated from ASV distribution.

190

191

Figure 3.18. The top five phyla (Acidobacteria, Actinobacteria, Bacteroidetes, Firmicutes and Proteobacteria) were distributed among all samples with the exception of Acidobacteria being the most clustered with rhizosphere samples. Principal coordinate analysis of bacterial communities of rhizosphere soil, root and stem samples and top five phyla. A: Distribution of bacterial communities by treatment, sample origin and sampling month. B: Distribution of top five phyla across samples. Samples that cluster close together share a greater similarity in composition.

191

Bibliography

Abbamondi, G. R., Tommonaro, G., Weyens, N., Thijs, S., Sillen, W., Gkorezis, P., Iodice, C., de Melo Rangel, W., Nicolaus, B., & Vangronsveld, J. (2016). Plant growth-promoting effects of rhizospheric and endophytic bacteria associated with different tomato cultivars and new tomato hybrids. Chemical and Biological Technologies in Agriculture, 3(1), 1. https://doi.org/10.1186/s40538- 015-0051-3

Acosta-Martínez, V., Bell, C. W., Morris, B. E. L., Zak, J., & Allen, V. G. (2010). Long-term soil microbial community and enzyme activity responses to an integrated cropping-livestock system in a semi-arid region☆. Agriculture, Ecosystems & Environment, 137(3–4), 231–240. https://doi.org/10.1016/j.agee.2010.02.008

Agnolucci, M., Palla, M., Cristani, C., Cavallo, N., Giovannetti, M., De Angelis, M., Gobbetti, M., & Minervini, F. (2019). Beneficial Plant Microorganisms Affect the Endophytic Bacterial Communities of Durum Wheat Roots as Detected by Different Molecular Approaches. Frontiers in Microbiology, 10. https://doi.org/10.3389/fmicb.2019.02500

Akbaba, M., & Ozaktan, H. (2018). Biocontrol of angular leaf spot disease and colonization of cucumber (Cucumis sativus L.) by endophytic bacteria. Egyptian Journal of Biological Pest Control, 28(1), 14. https://doi.org/10.1186/s41938- 017-0020-1

Akinsanya, M. A., Goh, J. K., Lim, S. P., & Ting, A. S. Y. (2015). Diversity, antimicrobial and antioxidant activities of culturable bacterial endophyte communities in Aloe vera. FEMS Microbiology Letters, 362(23). https://doi.org/10.1093/femsle/fnv184

Al-Kharousi, Z. S., Guizani, N., Al-Sadi, A. M., Al-Bulushi, I. M., & Shaharoona, B. (2016). Hiding in Fresh Fruits and Vegetables: Opportunistic Pathogens May Cross Geographical Barriers [Research Article]. International Journal of Microbiology; Hindawi. https://doi.org/10.1155/2016/4292417

192

Alaoui, A., Rogger, M., Peth, S., & Blöschl, G. (2018). Does soil compaction increase floods? A review. Journal of Hydrology, 557, 631–642. https://doi.org/10.1016/j.jhydrol.2017.12.052

Allard, S. M., Ottesen, A. R., & Micallef, S. A. (2020). Rain induces temporary shifts in epiphytic bacterial communities of cucumber and tomato fruit. Scientific Reports, 10(1), 1765. https://doi.org/10.1038/s41598-020-58671-7

Allard, S. M., Walsh, C. S., Wallis, A. E., Ottesen, A. R., Brown, E. W., & Micallef, S. A. (2016). Solanum lycopersicum (tomato) hosts robust phyllosphere and rhizosphere bacterial communities when grown in soil amended with various organic and synthetic fertilizers. Science of The Total Environment, 573, 555– 563. https://doi.org/10.1016/j.scitotenv.2016.08.157

Alphei, J., & Scheu, S. (1993). Effects of biocidal treatments on biological and nutritional properties of a mull-structured woodland soil. In L. Brussaard & M. J. Kooistra (Eds.), Soil Structure/Soil Biota Interrelationships (pp. 435–448). Elsevier. https://doi.org/10.1016/B978-0-444-81490-6.50035-2

Altschul, S. F., Gish, W., Miller, W., Myers, E. W., & Lipman, D. J. (1990). Basic local alignment search tool. Journal of Molecular Biology, 215(3), 403–410. https://doi.org/10.1016/S0022-2836(05)80360-2

Andersen, K. S., Kirkegaard, R. H., Karst, S. M., & Albertsen, M. (2018). ampvis2: An R package to analyse and visualise 16S rRNA amplicon data. BioRxiv, 299537. https://doi.org/10.1101/299537

Anderson, M. J. (2017). Permutational Multivariate Analysis of Variance (PERMANOVA). In Wiley StatsRef: Statistics Reference Online (pp. 1–15). American Cancer Society. https://doi.org/10.1002/9781118445112.stat07841

Antoniou, A., Tsolakidou, M.-D., Stringlis, I. A., & Pantelides, I. S. (2017). Rhizosphere Microbiome Recruited from a Suppressive Compost Improves Plant Fitness and Increases Protection against Vascular Wilt Pathogens of Tomato. Frontiers in Plant Science, 8. https://doi.org/10.3389/fpls.2017.02022

Araji, A. A., Abdo, Z. O., & Joyce, P. (2001). Efficient use of animal manure on cropland—Economic analysis. Bioresource Technology, 79(2), 179–191. https://doi.org/10.1016/s0960-8524(01)00042-6

Arfaoui, M., Vallance, J., Bruez, E., Rezgui, A., Melki, I., Chebil, S., Sadfi-Zouaoui, N., & Rey, P. (2019). Isolation, identification and in vitro characterization of grapevine rhizobacteria to control ochratoxigenic Aspergillus spp. On grapes.

193

Biological Control, 129, 201–211. https://doi.org/10.1016/j.biocontrol.2018.10.019

Armanhi, J. S. L., de Souza, R. S. C., Damasceno, N. de B., de Araújo, L. M., Imperial, J., & Arruda, P. (2018). A Community-Based Culture Collection for Targeting Novel Plant Growth-Promoting Bacteria from the Sugarcane Microbiome. Frontiers in Plant Science, 8. https://doi.org/10.3389/fpls.2017.02191

Aryantha, I. P., Cross, R., & Guest, D. I. (2000). Suppression of Phytophthora cinnamomi in Potting Mixes Amended with Uncomposted and Composted Animal Manures. Phytopathology®, 90(7), 775–782. https://doi.org/10.1094/PHYTO.2000.90.7.775

Bacon, C. W., & White, J. (2000). Microbial Endophytes. CRC Press.

Bai, Y., D’Aoust, F., Smith, D. L., & Driscoll, B. T. (2002). Isolation of plant-growth- promoting Bacillus strains from soybean root nodules. Canadian Journal of Microbiology, 48(3), 230–238. https://doi.org/10.1139/w02-014

Bailey, V. L., Fansler, S. J., Stegen, J. C., & McCue, L. A. (2013). Linking microbial community structure to β -glucosidic function in soil aggregates. The ISME Journal, 7(10), 2044–2053. https://doi.org/10.1038/ismej.2013.87

Beckers, B., Op De Beeck, M., Thijs, S., Truyens, S., Weyens, N., Boerjan, W., & Vangronsveld, J. (2016). Performance of 16s rDNA Primer Pairs in the Study of Rhizosphere and Endosphere Bacterial Microbiomes in Metabarcoding Studies. Frontiers in Microbiology, 7. https://doi.org/10.3389/fmicb.2016.00650

Beesley, C. A., Vanner, C. L., Helsel, L. O., Gee, J. E., & Hoffmaster, A. R. (2010). Identification and characterization of clinical Bacillus spp. Isolates phenotypically similar to Bacillus anthracis. FEMS Microbiology Letters, 313(1), 47–53. https://doi.org/10.1111/j.1574-6968.2010.02120.x

Bendaha, M. E. A., & Belaouni, H. A. (2019). Tomato growth and resistance promotion by Enterobacter hormaechei subsp. Steigerwaltii EB8D. Archives of Phytopathology and Plant Protection, 52(3–4), 318–332. https://doi.org/10.1080/03235408.2019.1620511

Bergna, A., Cernava, T., Rändler, M., Grosch, R., Zachow, C., & Berg, G. (2018). Tomato Seeds Preferably Transmit Plant Beneficial Endophytes. Phytobiomes Journal, 2(4), 183–193. https://doi.org/10.1094/PBIOMES-06-18-0029-R

194

Bergougnoux, V. (2014). The history of tomato: From domestication to biopharming. Biotechnology Advances, 32(1), 170–189. https://doi.org/10.1016/j.biotechadv.2013.11.003

Berns, A. E., Philipp, H., Narres, H.-D., Burauel, P., Vereecken, H., & Tappe, W. (2008). Effect of gamma-sterilization and autoclaving on soil organic matter structure as studied by solid state NMR, UV and fluorescence spectroscopy. European Journal of Soil Science, 59(3), 540–550. https://doi.org/10.1111/j.1365-2389.2008.01016.x

Bishop, J. P. (1995). Chickens: Improving small-scale production. Echo Technical Note, 10.

Bodenhausen, N., Bortfeld-Miller, M., Ackermann, M., & Vorholt, J. A. (2014). A Synthetic Community Approach Reveals Plant Genotypes Affecting the Phyllosphere Microbiota. PLoS Genetics, 10(4). https://doi.org/10.1371/journal.pgen.1004283

Bodenhausen, N., Horton, M. W., & Bergelson, J. (2013). Bacterial Communities Associated with the Leaves and the Roots of Arabidopsis thaliana. PLoS ONE, 8(2), e56329. https://doi.org/10.1371/journal.pone.0056329

Brady, C., Cleenwerck, I., Venter, S., Vancanneyt, M., Swings, J., & Coutinho, T. (2008). Phylogeny and identification of Pantoea species associated with plants, humans and the natural environment based on multilocus sequence analysis (MLSA). Systematic and Applied Microbiology, 31(6), 447–460. https://doi.org/10.1016/j.syapm.2008.09.004

Bulgari, D., Casati, P., Brusetti, L., Quaglino, F., Brasca, M., Daffonchio, D., & Bianco, P. A. (2009). Endophytic bacterial diversity in grapevine (Vitis vinifera L.) leaves described by 16S rRNA gene sequence analysis and length heterogeneity- PCR. The Journal of Microbiology, 47(4), 393–401. https://doi.org/10.1007/s12275-009-0082-1

Callahan, B. J., McMurdie, P. J., Rosen, M. J., Han, A. W., Johnson, A. J. A., & Holmes, S. P. (2016). DADA2: High-resolution sample inference from Illumina amplicon data. Nature Methods, 13(7), 581–583. https://doi.org/10.1038/nmeth.3869

Carper, D. L., Carrell, A. A., Kueppers, L. M., & Frank, A. C. (2018). Bacterial endophyte communities in Pinus flexilis are structured by host age, tissue type, and environmental factors. Plant and Soil; Dordrecht, 428(1–2), 335–352. http://dx.doi.org.proxy.lib.ohio-state.edu/10.1007/s11104-018-3682-x

195

Case, R. J., Boucher, Y., Dahllöf, I., Holmström, C., Doolittle, W. F., & Kjelleberg, S. (2007). Use of 16S rRNA and rpoB genes as molecular markers for microbial ecology studies. Applied and Environmental Microbiology, 73(1), 278–288. https://doi.org/10.1128/AEM.01177-06

CFAES. (2020). College of Food, Agricultural, and Environmental Sciences Weather System: Wooster Station. CFAES Weather System, The Ohio State University. https://www.oardc.ohio-state.edu/weather1/

Ch’ng, H. Y., Ahmed, O. H., Kassim, S., & Majid, N. M. A. (2013). Co-composting of pineapple leaves and chicken manure slurry. International Journal Of Recycling of Organic Waste in Agriculture, 2(1), 23. https://doi.org/10.1186/2251-7715-2- 23

Chao, A., Chazdon, R. L., Colwell, R. K., & Shen, T.-J. (2006). Abundance-based similarity indices and their estimation when there are unseen species in samples. Biometrics, 62(2), 361–371. https://doi.org/10.1111/j.1541-0420.2005.00489.x

Chaparro, J. M., Sheflin, A. M., Manter, D. K., & Vivanco, J. M. (2012). Manipulating the soil microbiome to increase soil health and plant fertility. Biology and Fertility of Soils, 48(5), 489–499. https://doi.org/10.1007/s00374-012-0691-4

Chastain, J. P., Camberato, J. J., & Skewes, P. (2001). Poultry Manure Production and Nutrient Content. 17.

Chaudhry, V., Rehman, A., Mishra, A., Chauhan, P. S., & Nautiyal, C. S. (2012). Changes in Bacterial Community Structure of Agricultural Land Due to Long- Term Organic and Chemical Amendments. Microbial Ecology, 64(2), 450–460. https://doi.org/10.1007/s00248-012-0025-y

Chauhan, A., Guleria, S., Walia, A., Mahajan, R., Verma, S., & Shirkot, C. K. (2014). Isolation and characterization of Bacillus sp. With their effect on growth of tomato seedlings. Indian Journal of Agricultural Biochemistry, 27(2), 193–201.

Chelius, M. K., & Triplett, E. W. (2000). Immunolocalization of Dinitrogenase Reductase Produced by Klebsiella pneumoniae in Association with Zea mays L. Applied and Environmental Microbiology, 66(2), 783–787. https://doi.org/10.1128/AEM.66.2.783-787.2000

Chelius, M. K., & Triplett, E. W. (2001). The Diversity of Archaea and Bacteria in Association with the Roots of Zea mays L. Microbial Ecology, 41(3), 252–263. https://doi.org/10.1007/s002480000087

196

Chellemi, D. O., Rosskopf, E. N., & Kokalis-Burelle, N. (2013). The Effect of Transitional Organic Production Practices on Soilborne Pests of Tomato in a Simulated Microplot Study. Phytopathology, 103(8), 792–801. https://doi.org/10.1094/PHYTO-09-12-0243-R

Chen, C., Chen, H. Y. H., Chen, X., & Huang, Z. (2019). Meta-analysis shows positive effects of plant diversity on microbial biomass and respiration. Nature Communications, 10(1), 1–10. https://doi.org/10.1038/s41467-019-09258-y

Chen, Z., & Jiang, X. (2014). Microbiological Safety of Chicken Litter or Chicken Litter-Based Organic Fertilizers: A Review. Agriculture, 4(1), 1–29. https://doi.org/10.3390/agriculture4010001

Cheng, Z., Lei, S., Li, Y., Huang, W., Ma, R., Xiong, J., Zhang, T., Jin, L., Haq, H. ul, Xu, X., & Tian, B. (2020). Revealing the Variation and Stability of Bacterial Communities in Tomato Rhizosphere Microbiota. Microorganisms, 8(2). https://doi.org/10.3390/microorganisms8020170

Chihaoui, S.-A., Trabelsi, D., Jdey, A., Mhadhbi, H., & Mhamdi, R. (2015). Inoculation of Phaseolus vulgaris with the nodule-endophyte Agrobacterium sp. 10C2 affects richness and structure of rhizosphere bacterial communities and enhances nodulation and growth. Archives of Microbiology, 197(6), 805–813. https://doi.org/10.1007/s00203-015-1118-z

Cohen, A. C., Travaglia, C. N., Bottini, R., & Piccoli, P. N. (2009). Participation of abscisic acid and gibberellins produced by endophytic Azospirillum in the alleviation of drought effects in maize. Botany, 87(5), 455–462. https://doi.org/10.1139/B09-023

Compant, S., Clément, C., & Sessitsch, A. (2010). Plant growth-promoting bacteria in the rhizo- and endosphere of plants: Their role, colonization, mechanisms involved and prospects for utilization. Soil Biology and Biochemistry, 42(5), 669–678. https://doi.org/10.1016/j.soilbio.2009.11.024

Compant, S., Mitter, B., Colli-Mull, J. G., Gangl, H., & Sessitsch, A. (2011). Endophytes of Grapevine Flowers, Berries, and Seeds: Identification of Cultivable Bacteria, Comparison with Other Plant Parts, and Visualization of Niches of Colonization. Microbial Ecology, 62(1), 188–197. https://doi.org/10.1007/s00248-011-9883-y

Compant, S., Reiter, B., Sessitsch, A., Nowak, J., Clément, C., & Barka, E. A. (2005). Endophytic Colonization of Vitis vinifera L. by Plant Growth-Promoting Bacterium Burkholderia sp. Strain PsJN. Applied and Environmental

197

Microbiology, 71(4), 1685–1693. https://doi.org/10.1128/AEM.71.4.1685- 1693.2005

Conn, K. L., & Lazarovits, G. (1999). Impact of animal manures on verticillium wilt, potato scab, and soil microbial populations. Canadian Journal of Plant Pathology, 21(1), 81–92. https://doi.org/10.1080/07060661.1999.10600089

Connon, S. A., & Giovannoni, S. J. (2002). High-Throughput Methods for Culturing Microorganisms in Very-Low-Nutrient Media Yield Diverse New Marine Isolates. Applied and Environmental Microbiology, 68(8), 3878–3885. https://doi.org/10.1128/AEM.68.8.3878-3885.2002

Copeland, J. K., Yuan, L., Layeghifard, M., Wang, P. W., & Guttman, D. S. (2015). Seasonal Community Succession of the Phyllosphere Microbiome. Molecular Plant-Microbe Interactions®, 28(3), 274–285. https://doi.org/10.1094/MPMI- 10-14-0331-FI

Cordero, J., de Freitas, J. R., & Germida, J. J. (2019). Bacterial microbiome associated with the rhizosphere and root interior of crops in Saskatchewan, Canada. Canadian Journal of Microbiology, 66(1), 71–85. https://doi.org/10.1139/cjm- 2019-0330

Cordovez, V., Schop, S., Hordijk, K., Boulois, H. D. de, Coppens, F., Hanssen, I., Raaijmakers, J. M., & Carrión, V. J. (2018). Priming of Plant Growth Promotion by Volatiles of Root-Associated Microbacterium spp. Applied and Environmental Microbiology, 84(22). https://doi.org/10.1128/AEM.01865-18

Coutinho, T. A., & Venter, S. N. (2009). Pantoea ananatis: An unconventional plant pathogen. Molecular Plant Pathology, 10(3), 325–335. https://doi.org/10.1111/j.1364-3703.2009.00542.x

Cramer, G. R., & Quarrie, S. A. (2002). Abscisic acid is correlated with the leaf growth inhibition of four genotypes of maize differing in their response to salinity. Functional Plant Biology, 29(1), 111–115. https://doi.org/10.1071/pp01131

Cressman, M. D., Yu, Z., Nelson, M. C., Moeller, S. J., Lilburn, M. S., & Zerby, H. N. (2010). Interrelations between the Microbiotas in the Litter and in the Intestines of Commercial Broiler Chickens. Applied and Environmental Microbiology, 76(19), 6572–6582. https://doi.org/10.1128/AEM.00180-10

Cruz, A. T., Cazacu, A. C., & Allen, C. H. (2007). Pantoea agglomerans, a Plant Pathogen Causing Human Disease. Journal of Clinical Microbiology, 45(6), 1989–1992. https://doi.org/10.1128/JCM.00632-07

198

Culman, S. W., Snapp, S. S., Freeman, M. A., Schipanski, M. E., Beniston, J., Lal, R., Drinkwater, L. E., Franzluebbers, A. J., Glover, J. D., Grandy, A. S., Lee, J., Six, J., Maul, J. E., Mirksy, S. B., Spargo, J. T., & Wander, M. M. (2012). Permanganate Oxidizable Carbon Reflects a Processed Soil Fraction that is Sensitive to Management. Soil Science Society of America Journal, 76(2), 494– 504. https://doi.org/10.2136/sssaj2011.0286

Davis, N. M., Proctor, D. M., Holmes, S. P., Relman, D. A., & Callahan, B. J. (2018). Simple statistical identification and removal of contaminant sequences in marker-gene and metagenomics data. Microbiome, 6(1), 226. https://doi.org/10.1186/s40168-018-0605-2 de Abreu, C. S., Figueiredo, J. E. F., Oliveira, C. A., dos Santos, V. L., Gomes, E. A., Ribeiro, V. P., Barros, B. A., Lana, U. G. P., & Marriel, I. E. (2017). Maize endophytic bacteria as mineral phosphate solubilizers. Genetics and Molecular Research, 16(1). https://doi.org/10.4238/gmr16019294

De Clerck, E., & De Vos, P. (2004). Genotypic diversity among Bacillus licheniformis strains from various sources. FEMS Microbiology Letters, 231(1), 91–98. https://doi.org/10.1016/S0378-1097(03)00935-2 de Faccio Carvalho, P. C., Anghinoni, I., de Moraes, A., de Souza, E. D., Sulc, R. M., Lang, C. R., Flores, J. P. C., Terra Lopes, M. L., da Silva, J. L. S., Conte, O., de Lima Wesp, C., Levien, R., Fontaneli, R. S., & Bayer, C. (2010). Managing grazing animals to achieve nutrient cycling and soil improvement in no-till integrated systems. Nutrient Cycling in Agroecosystems, 88(2), 259–273. https://doi.org/10.1007/s10705-010-9360-x de Mendiburu, F. (2020). agricolae: Statistical Procedures for Agricultural Research. (Version R package version 1.3-2) [Computer software]. https://CRAN.R- project.org/package=agricolae

DEFRA. (2001). The welfare of hens in free range systems. Department for Environment, Food & Rural Affairs, U.K.

Delmotte, N., Knief, C., Chaffron, S., Innerebner, G., Roschitzki, B., Schlapbach, R., Mering, C. von, & Vorholt, J. A. (2009). Community proteogenomics reveals insights into the physiology of phyllosphere bacteria. Proceedings of the National Academy of Sciences, 106(38), 16428–16433. https://doi.org/10.1073/pnas.0905240106

Demir, K., Sahin, O., Kadioglu, Y. K., Pilbeam, D. J., & Gunes, A. (2010). Essential and non-essential element composition of tomato plants fertilized with poultry

199

manure. Scientia Horticulturae, 127(1), 16–22. https://doi.org/10.1016/j.scienta.2010.08.009

Deng, W., Zhang, A., Chen, S., He, X., Jin, L., Yu, X., Yang, S., Li, B., Fan, L., Ji, L., Pan, X., & Zou, L. (2020). Heavy metals, antibiotics and nutrients affect the bacterial community and resistance genes in chicken manure composting and fertilized soil. Journal of Environmental Management, 257, 109980. https://doi.org/10.1016/j.jenvman.2019.109980

Deng, Z.-S., Zhang, B.-C., Qi, X.-Y., Sun, Z.-H., He, X.-L., Liu, Y.-Z., Li, J., Chen, K.- K., & Lin, Z.-X. (2019). Root-Associated Endophytic Bacterial Community Composition of Pennisetum sinese from Four Representative Provinces in China. Microorganisms, 7(2). https://doi.org/10.3390/microorganisms7020047

Devendra, C., & Thomas, D. (2002). Crop–animal interactions in mixed farming systems in Asia. Agricultural Systems, 71(1–2), 27–40. https://doi.org/10.1016/S0308-521X(01)00034-8

Dikinya, O., & Mufwanzala, N. (2010). Chicken manure-enhanced soil fertility and productivity: Effects of application rates. Journal of Soil Science and Environmental Management, 1(3), 46–54.

Ding, G.-C., Radl, V., Schloter-Hai, B., Jechalke, S., Heuer, H., Smalla, K., & Schloter, M. (2014). Dynamics of Soil Bacterial Communities in Response to Repeated Application of Manure Containing Sulfadiazine. PLOS ONE, 9(3), e92958. https://doi.org/10.1371/journal.pone.0092958

Ding, T., & Melcher, U. (2016). Influences of Plant Species, Season and Location on Leaf Endophytic Bacterial Communities of Non-Cultivated Plants. PLoS ONE, 11(3). https://doi.org/10.1371/journal.pone.0150895

Ding, T., Palmer, M. W., & Melcher, U. (2013). Community terminal restriction fragment length polymorphisms reveal insights into the diversity and dynamics of leaf endophytic bacteria. BMC Microbiology, 13(1), 1. https://doi.org/10.1186/1471-2180-13-1

Dong, C.-J., Wang, L.-L., Li, Q., & Shang, Q.-M. (2019). Bacterial communities in the rhizosphere, phyllosphere and endosphere of tomato plants. PLoS ONE, 14(11). https://doi.org/10.1371/journal.pone.0223847

Duijff, B. J., Gianinazzi‐Pearson, V., & Lemanceau, P. (1997). Involvement of the outer membrane lipopolysaccharides in the endophytic colonization of tomato roots by biocontrol Pseudomonas fluorescens strain WCS417r. New Phytologist, 135(2), 325–334. https://doi.org/10.1046/j.1469-8137.1997.00646.x 200

Dursun, A., Ekinci, M., & Dönmez, M. F. (2010). Effects of foliar application of plant growth promoting bacterium on chemical contents, yield and growth of tomato (Lycopersicon esculentum L.) and cucumber (Cucumis sativus L.). Pakistan Journal of Botany, 42, 3349–3356.

Edwards, J. A., Santos-Medellín, C. M., Liechty, Z. S., Nguyen, B., Lurie, E., Eason, S., Phillips, G., & Sundaresan, V. (2018). Compositional shifts in root-associated bacterial and archaeal microbiota track the plant life cycle in field-grown rice. PLOS Biology, 16(2), e2003862. https://doi.org/10.1371/journal.pbio.2003862

Enya, J., Koitabashi, M., Shinohara, H., Yoshida, S., Tsukiboshi, T., Negishi, H., Suyama, K., & Tsushima, S. (2007). Phylogenetic Diversities of Dominant Culturable Bacillus, Pseudomonas and Pantoea Species on Tomato Leaves and Their Possibility as Biological Control Agents. Journal of Phytopathology, 155(7–8), 446–453. https://doi.org/10.1111/j.1439-0434.2007.01256.x

Enya, J., Shinohara, H., Yoshida, S., Tsukiboshi, T., Negishi, H., Suyama, K., & Tsushima, S. (2007). Culturable Leaf-Associated Bacteria on Tomato Plants and Their Potential as Biological Control Agents. Microbial Ecology, 53(4), 524– 536. https://doi.org/10.1007/s00248-006-9085-1

Eribo, B., & Ashenafi, M. (2003). Behavior of Escherichia coli O157:H7 in tomato and processed tomato products. Food Research International, 36(8), 823–830. https://doi.org/10.1016/S0963-9969(03)00077-2

Faith, D. P. (2018). Phylogenetic Diversity and Conservation Evaluation: Perspectives on Multiple Values, Indices, and Scales of Application. In R. A. Scherson & D. P. Faith (Eds.), Phylogenetic Diversity (pp. 1–26). Springer International Publishing. https://doi.org/10.1007/978-3-319-93145-6_1

Fan, F., Yin, C., Tang, Y., Li, Z., Song, A., Wakelin, S. A., Zou, J., & Liang, Y. (2014). Probing potential microbial coupling of carbon and nitrogen cycling during decomposition of maize residue by 13C-DNA-SIP. Soil Biology and Biochemistry, 70, 12–21. https://doi.org/10.1016/j.soilbio.2013.12.002

FAO. (2018). FAOSTAT: Food and agriculture data. Food and Agriculture Organization of the United Nations. http://www.fao.org/faostat/en/

Fierer, N., Bradford, M. A., & Jackson, R. B. (2007). Toward an ecological classification of soil bacteria. Ecology, 88(6), 1354–1364. https://doi.org/10.1890/05-1839

201

Fox, J., & Weisberg, S. (2019). An {R} Companion to Applied Regression (Third). Sage. https://socialsciences.mcmaster.ca/jfox/Books/Companion/

Franzluebbers, A. J., Haney, R. L., Honeycutt, C. W., Schomberg, H. H., & Hons, F. M. (2000). Flush of Carbon Dioxide Following Rewetting of Dried Soil Relates to Active Organic Pools. Soil Science Society of America Journal, 64(2), 613–623. https://doi.org/10.2136/sssaj2000.642613x

Gao, Y., Tian, Y., Liang, X., & Gao, L. (2015). Effects of single-root-grafting, double- root-grafting and compost application on microbial properties of rhizosphere soils in Chinese protected cucumber (Cucumis sativus L.) production systems. Scientia Horticulturae, 186, 190–200. https://doi.org/10.1016/j.scienta.2015.02.026

Gardner, J. C., Faulkner, D. B., & Hargrove, W. L. (1991). Use of cover crops with integrated crop-livestock production systems. Cover Crops for Clean Water. Soil and Water Conserv. Soc., Ankeny, IA, 185–191.

Gardner, J. M., Feldman, A. W., & Zablotowicz, R. M. (1982). Identity and behavior of xylem-residing bacteria in rough lemon roots of Florida citrus trees. Applied and Environmental Microbiology, 43(6), 1335–1342.

Gdanetz, K., & Trail, F. (2017). The Wheat Microbiome Under Four Management Strategies, and Potential for Endophytes in Disease Protection. Phytobiomes Journal, 1(3), 158–168. https://doi.org/10.1094/PBIOMES-05-17-0023-R

Genomics Workbench version 20.0. (2020). https://digitalinsights.qiagen.com

Gerberastraat, B. V., & Aalsmeer, R. A. (2001). 4Peaks. Nucleobytes. https://nucleobytes.com/4peaks/index.html

Ghyselinck, J., Pfeiffer, S., Heylen, K., Sessitsch, A., & De Vos, P. (2013). The Effect of Primer Choice and Short Read Sequences on the Outcome of 16S rRNA Gene Based Diversity Studies. PLoS ONE, 8(8), e71360. https://doi.org/10.1371/journal.pone.0071360

Glaeser, S. P., & Kämpfer, P. (2015). Multilocus sequence analysis (MLSA) in prokaryotic taxonomy. Systematic and Applied Microbiology, 38(4), 237–245. https://doi.org/10.1016/j.syapm.2015.03.007

Glaeser, S. P., Falsen, E., Busse, H.-J., & Kämpfer, P. (2013). Paenibacillus vulneris sp. Nov., isolated from a necrotic wound. International Journal of Systematic and Evolutionary Microbiology, 63(Pt 2), 777–782. https://doi.org/10.1099/ijs.0.041210-0 202

Glassing, A., Dowd, S. E., Galandiuk, S., Davis, B., & Chiodini, R. J. (2016). Inherent bacterial DNA contamination of extraction and sequencing reagents may affect interpretation of microbiota in low bacterial biomass samples. Gut Pathogens, 8(1), 24. https://doi.org/10.1186/s13099-016-0103-7

Gomez, E., Ferreras, L., & Toresani, S. (2006). Soil bacterial functional diversity as influenced by organic amendment application. Bioresource Technology, 97(13), 1484–1489. https://doi.org/10.1016/j.biortech.2005.06.021

Gonçalves, R., Balbi-Peña, M., Soman, J., Maringoni, A., Taghouti, G., Fischer-Le Saux, M., & Portier, P. (2018). Genetic diversity of Curtobacterium flaccumfaciens revealed by multilocus sequence analysis. European Journal of Plant Pathology. https://doi.org/10.1007/s10658-018-01648-0

Gong, J., Si, W., Forster, R. J., Huang, R., Yu, H., Yin, Y., Yang, C., & Han, Y. (2007). 16S rRNA gene-based analysis of mucosa-associated bacterial community and phylogeny in the chicken gastrointestinal tracts: From crops to ceca. FEMS Microbiology Ecology, 59(1), 147–157. https://doi.org/10.1111/j.1574- 6941.2006.00193.x

Gupta, P., Kumar, V., Usmani, Z., Rani, R., Chandra, A., & Gupta, V. K. (2020). Implications of plant growth promoting Klebsiella sp. CPSB4 and Enterobacter sp. CPSB49 in luxuriant growth of tomato plants under chromium stress. Chemosphere, 240, 124944. https://doi.org/10.1016/j.chemosphere.2019.124944

Gupta, V., Rai, P. K., & Risam, K. S. (2012). Integrated crop-livestock farming systems: A strategy for resource conservation and environmental sustainability. Indian Research Journal of Extension Education, Special Issue, 2, 49–54.

Hacquard, S., Garrido-Oter, R., González, A., Spaepen, S., Ackermann, G., Lebeis, S., McHardy, A. C., Dangl, J. L., Knight, R., Ley, R., & Schulze-Lefert, P. (2015). Microbiota and Host Nutrition across Plant and Animal Kingdoms. Cell Host & Microbe, 17(5), 603–616. https://doi.org/10.1016/j.chom.2015.04.009

Hallmann, J., Quadt-Hallmann, A., Mahaffee, W. F., & Kloepper, J. W. (1997). Bacterial endophytes in agricultural crops. Canadian Journal of Microbiology, 43(10), 895–914. https://doi.org/10.1139/m97-131

Hallmann, Johannes, Berg, G., & Schulz, B. (2006). Isolation Procedures for Endophytic Microorganisms. In B. J. E. Schulz, C. J. C. Boyle, & T. N. Sieber (Eds.), Microbial Root Endophytes (pp. 299–319). Springer. https://doi.org/10.1007/3-540-33526-9_17

203

Hardoim, Pablo R., Hardoim, C. C. P., Overbeek, L. S. van, & Elsas, J. D. van. (2012). Dynamics of Seed-Borne Rice Endophytes on Early Plant Growth Stages. PLOS ONE, 7(2), e30438. https://doi.org/10.1371/journal.pone.0030438

Hardoim, Pablo R., van Overbeek, L. S., Berg, G., Pirttilä, A. M., Compant, S., Campisano, A., Döring, M., & Sessitsch, A. (2015). The Hidden World within Plants: Ecological and Evolutionary Considerations for Defining Functioning of Microbial Endophytes. Microbiology and Molecular Biology Reviews, 79(3), 293–320. https://doi.org/10.1128/MMBR.00050-14

Hardoim, Pablo Rodrigo, Nazir, R., Sessitsch, A., Elhottová, D., Korenblum, E., van Overbeek, L. S., & van Elsas, J. D. (2013). The new species Enterobacter oryziphilus sp. Nov. And Enterobacter oryzendophyticus sp. Nov. Are key inhabitants of the endosphere of rice. BMC Microbiology, 13, 164. https://doi.org/10.1186/1471-2180-13-164

Hartman, K., van der Heijden, M. G. A., Wittwer, R. A., Banerjee, S., Walser, J.-C., & Schlaeppi, K. (2018). Cropping practices manipulate abundance patterns of root and soil microbiome members paving the way to smart farming. Microbiome, 6(1), 14. https://doi.org/10.1186/s40168-017-0389-9

Hartmann, M., Frey, B., Mayer, J., Mäder, P., & Widmer, F. (2015). Distinct soil microbial diversity under long-term organic and conventional farming. The ISME Journal, 9(5), 1177–1194. https://doi.org/10.1038/ismej.2014.210

He, T., & Cramer, G. R. (1996). Abscisic acid concentrations are correlated with leaf area reductions in two salt-stressed rapid-cycling Brassica species. Plant and Soil, 179(1), 25–33. https://doi.org/10.1007/BF00011639

Herbold, C. W., Pelikan, C., Kuzyk, O., Hausmann, B., Angel, R., Berry, D., & Loy, A. (2015). A flexible and economical barcoding approach for highly multiplexed amplicon sequencing of diverse target genes. Frontiers in Microbiology, 6. https://doi.org/10.3389/fmicb.2015.00731

Herrera-Medina, M. J., Steinkellner, S., Vierheilig, H., Ocampo Bote, J. A., & García Garrido, J. M. (2007). Abscisic acid determines arbuscule development and functionality in the tomato arbuscular mycorrhiza. New Phytologist, 175(3), 554–564. https://doi.org/10.1111/j.1469-8137.2007.02107.x

Hilimire, K., Gliessman, S., & Muramoto, J. (2013). Soil fertility and crop growth under poultry/crop integration. Renewable Agriculture and Food Systems, 28, 173– 182. https://doi.org/10.1017/S174217051200021X

204

Ho, A., Di Lonardo, D. P., & Bodelier, P. L. E. (2017). Revisiting life strategy concepts in environmental microbial ecology. FEMS Microbiology Ecology, 93(3). https://doi.org/10.1093/femsec/fix006

Hoffmann, H., & Roggenkamp, A. (2003). Population Genetics of the Nomenspecies Enterobacter cloacae. Appl. Environ. Microbiol., 69(9), 5306–5318. https://doi.org/10.1128/AEM.69.9.5306-5318.2003

Hole, D. G., Perkins, A. J., Wilson, J. D., Alexander, I. H., Grice, P. V., & Evans, A. D. (2005). Does organic farming benefit biodiversity? Biological Conservation, 122(1), 113–130. https://doi.org/10.1016/j.biocon.2004.07.018

Hornung, B. V. H., Zwittink, R. D., & Kuijper, E. J. (2019). Issues and current standards of controls in microbiome research. FEMS Microbiology Ecology, 95(5). https://doi.org/10.1093/femsec/fiz045

Horton, M. W., Bodenhausen, N., Beilsmith, K., Meng, D., Subramanian, S., Vetter, M. M., Vilhjálmsson, B. J., Gordon, J. I., & Bergelson, J. (2015). Genome-wide association study of Arabidopsis thaliana’s leaf microbial community. Nature Communications, 5(1), 1–7.

Hu, H. Q., Li, X. S., & He, H. (2010). Characterization of an antimicrobial material from a newly isolated Bacillus amyloliquefaciens from mangrove for biocontrol of Capsicum bacterial wilt. Biological Control, 54(3), 359–365. https://doi.org/10.1016/j.biocontrol.2010.06.015

Hurisso, T. T., Moebius‐Clune, D. J., Culman, S. W., Moebius‐Clune, B. N., Thies, J. E., & Es, H. M. van. (2018). Soil Protein as a Rapid Soil Health Indicator of Potentially Available Organic Nitrogen. Agricultural & Environmental Letters, 3(1), 180006. https://doi.org/10.2134/ael2018.02.0006

Ismail, A.-S., Eissa, A., El-Beltagy, A., & Abou Hadid, A. (1996). Tomato growth in calcareous soils in relation to forms and levels of some macro- and micronutrients. Acta Horticulturae, 434, 85–94. https://doi.org/10.17660/ActaHortic.1996.434.9

Jack, C. N., Wozniak, K. J., Porter, S. S., & Friesen, M. L. (2019). Rhizobia protect their legume hosts against soil-borne microbial antagonists in a host-genotype- dependent manner. Rhizosphere, 9, 47–55. https://doi.org/10.1016/j.rhisph.2018.11.005

Jeong, J.-J., Lee, D. W., Park, B., Sang, M. K., Choi, I.-G., & Kim, K. D. (2017). Chryseobacterium cucumeris sp. Nov., an endophyte isolated from cucumber (Cucumis sativus L.) root, and emended description of Chryseobacterium 205

arthrosphaerae. International Journal of Systematic and Evolutionary Microbiology, 67(3), 610–616. https://doi.org/10.1099/ijsem.0.001670

Jiao, N. (2013). Developing a Novel Approach of rpoB Gene as a Powerful Biomarker for the Environmental Microbial Diversity. Geomicrobiology Journal, 30(2), 108–119. https://doi.org/10.1080/01490451.2011.653090

Jin, X., Shi, Y., Wu, F., Pan, K., Zhou, X., Jin, X., Shi, Y., Wu, F., Pan, K., & Zhou, X. (2020). Intercropping of wheat changed cucumber rhizosphere bacterial community composition and inhibited cucumber Fusarium wilt disease. Scientia Agricola, 77(5). https://doi.org/10.1590/1678-992x-2019-0005

Kawasaki, A., Donn, S., Ryan, P. R., Mathesius, U., Devilla, R., Jones, A., & Watt, M. (2016). Microbiome and Exudates of the Root and Rhizosphere of Brachypodium distachyon, a Model for Wheat. PloS One, 11(10), e0164533. https://doi.org/10.1371/journal.pone.0164533

Kefi, A., Slimene, I. B., Karkouch, I., Rihouey, C., Azaeiz, S., Bejaoui, M., Belaid, R., Cosette, P., Jouenne, T., & Limam, F. (2015). Characterization of endophytic Bacillus strains from tomato plants (Lycopersicon esculentum) displaying antifungal activity against Botrytis cinerea Pers. World Journal of Microbiology and Biotechnology, 31(12), 1967–1976. https://doi.org/10.1007/s11274-015- 1943-x

Khalaf, E. M., & Raizada, M. N. (2016). Taxonomic and functional diversity of cultured seed associated microbes of the cucurbit family. BMC Microbiology, 16(1), 131. https://doi.org/10.1186/s12866-016-0743-2

Khan, A. L., Waqas, M., Kang, S.-M., Al-Harrasi, A., Hussain, J., Al-Rawahi, A., Al- Khiziri, S., Ullah, I., Ali, L., Jung, H.-Y., & Lee, I.-J. (2014). Bacterial endophyte Sphingomonas sp. LK11 produces gibberellins and IAA and promotes tomato plant growth. Journal of Microbiology, 52(8), 689–695. https://doi.org/10.1007/s12275-014-4002-7

Kim, M., Morrison, M., & Yu, Z. (2011). Evaluation of different partial 16S rRNA gene sequence regions for phylogenetic analysis of microbiomes. Journal of Microbiological Methods, 84(1), 81–87. https://doi.org/10.1016/j.mimet.2010.10.020

Knief, C., Delmotte, N., Chaffron, S., Stark, M., Innerebner, G., Wassmann, R., Mering, C. von, & Vorholt, J. A. (2012). Metaproteogenomic analysis of microbial communities in the phyllosphere and rhizosphere of rice. The ISME Journal, 6(7), 1378–1390. https://doi.org/10.1038/ismej.2011.192

206

Kumar, S., Stecher, G., & Tamura, K. (2016). MEGA7: Molecular Evolutionary Genetics Analysis Version 7.0 for Bigger Datasets. Molecular Biology and Evolution, 33(7), 1870–1874. https://doi.org/10.1093/molbev/msw054

Kunova, A., Bonaldi, M., Saracchi, M., Pizzatti, C., Chen, X., & Cortesi, P. (2016). Selection of Streptomyces against soil borne fungal pathogens by a standardized dual culture assay and evaluation of their effects on seed germination and plant growth. BMC Microbiology, 16(1), 272. https://doi.org/10.1186/s12866-016- 0886-1

Kwak, M.-J., Kong, H. G., Choi, K., Kwon, S.-K., Song, J. Y., Lee, J., Lee, P. A., Choi, S. Y., Seo, M., Lee, H. J., Jung, E. J., Park, H., Roy, N., Kim, H., Lee, M. M., Rubin, E. M., Lee, S.-W., & Kim, J. F. (2018). Rhizosphere microbiome structure alters to enable wilt resistance in tomato. Nature Biotechnology, 36(11), 1100–1109. https://doi.org/10.1038/nbt.4232

Kyakuwaire, M., Olupot, G., Amoding, A., Nkedi-Kizza, P., & Ateenyi Basamba, T. (2019). How Safe is Chicken Litter for Land Application as an Organic Fertilizer?: A Review. International Journal of Environmental Research and Public Health, 16(19). https://doi.org/10.3390/ijerph16193521

Lai, L., Zhao, X., Jiang, L., Wang, Y., Luo, L., Zheng, Y., Chen, X., & Rimmington, G. M. (2012). Soil Respiration in Different Agricultural and Natural Ecosystems in an Arid Region. PLOS ONE, 7(10), e48011. https://doi.org/10.1371/journal.pone.0048011

Lebuhn, M., Achouak, W., Schloter, M., Berge, O., Meier, H., Barakat, M., Hartmann, A., & Heulin, T. (2000). Taxonomic characterization of Ochrobactrum sp. Isolates from soil samples and wheat roots, and description of Ochrobactrum tritici sp. Nov. And Ochrobactrum grignonense sp. Nov. International Journal of Systematic and Evolutionary Microbiology, 50(6), 2207–2223. https://doi.org/10.1099/00207713-50-6-2207

Lee, S. A., Park, J., Chu, B., Kim, J. M., Joa, J.-H., Sang, M. K., Song, J., & Weon, H.- Y. (2016). Comparative analysis of bacterial diversity in the rhizosphere of tomato by culture-dependent and -independent approaches. Journal of Microbiology, 54(12), 823–831. https://doi.org/10.1007/s12275-016-6410-3

Leite, H. A. C., Silva, A. B., Gomes, F. P., Gramacho, K. P., Faria, J. C., de Souza, J. T., & Loguercio, L. L. (2013). Bacillus subtilis and Enterobacter cloacae endophytes from healthy Theobroma cacao L. trees can systemically colonize seedlings and promote growth. Applied Microbiology and Biotechnology, 97(6), 2639–2651. https://doi.org/10.1007/s00253-012-4574-2

207

Lenth, R. V. (2016). Least-Squares Means: The R Package lsmeans. Journal of Statistical Software, 69(1), 1–33. https://doi.org/10.18637/jss.v069.i01

Lin, L., Guo, W., Xing, Y., Zhang, X., Li, Z., Hu, C., Li, S., Li, Y., & An, Q. (2012). The actinobacterium Microbacterium sp. 16SH accepts pBBR1-based pPROBE vectors, forms biofilms, invades roots, and fixes N2 associated with micropropagated sugarcane plants. Applied Microbiology and Biotechnology, 93(3), 1185–1195. https://doi.org/10.1007/s00253-011-3618-3

Lipiec, J., & Hatano, R. (2003). Quantification of compaction effects on soil physical properties and crop growth. Geoderma, 116(1), 107–136. https://doi.org/10.1016/S0016-7061(03)00097-1

Liu, L., Chen, H., Cai, P., Liang, W., & Huang, Q. (2009). Immobilization and phytotoxicity of Cd in contaminated soil amended with chicken manure compost. Journal of Hazardous Materials, 163(2–3), 563–567. https://doi.org/10.1016/j.jhazmat.2008.07.004

Lopez-Fernandez, S., Compant, S., Vrhovsek, U., Bianchedi, P. L., Sessitsch, A., Pertot, I., & Campisano, A. (2016). Grapevine colonization by endophytic bacteria shifts secondary metabolism and suggests activation of defense pathways. Plant and Soil, 405(1), 155–175. https://doi.org/10.1007/s11104-015-2631-1

López, S., Pastorino, G., Franco, M., Medina, R., Lucentini, C., & Balatti, P. (2018). Microbial Endophytes that Live within the Seeds of Two Tomato Hybrids Cultivated in Argentina. Agronomy, 8, 136. https://doi.org/10.3390/agronomy8080136

Love, M. I., Huber, W., & Anders, S. (2014). Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biology, 15(12), 550. https://doi.org/10.1186/s13059-014-0550-8

Lozupone, C., & Knight, R. (2005). UniFrac: A New Phylogenetic Method for Comparing Microbial Communities. Applied and Environmental Microbiology, 71(12), 8228–8235. https://doi.org/10.1128/AEM.71.12.8228-8235.2005

Lucaciu, R., Pelikan, C., Gerner, S. M., Zioutis, C., Köstlbacher, S., Marx, H., Herbold, C. W., Schmidt, H., & Rattei, T. (2019). A Bioinformatics Guide to Plant Microbiome Analysis. Frontiers in Plant Science, 10. https://doi.org/10.3389/fpls.2019.01313

Luo, J., Wyatt, J., van der Weerden, T. J., Thomas, S. M., de Klein, C. A. M., Li, Y., Rollo, M., Lindsey, S., Ledgard, S. F., Li, J., Ding, W., Qin, S., Zhang, N., Bolan, N., Kirkham, M. B., Bai, Z., Ma, L., Zhang, X., Wang, H., … Rys, G. 208

(2017). Chapter Five—Potential Hotspot Areas of Nitrous Oxide Emissions From Grazed Pastoral Dairy Farm Systems. In D. L. Sparks (Ed.), Advances in Agronomy (Vol. 145, pp. 205–268). Academic Press. https://doi.org/10.1016/bs.agron.2017.05.006

Luo, T., Ou‐Yang, X.-Q., Yang, L.-T., Li, Y.-R., Song, X.-P., Zhang, G.-M., Gao, Y.-J., Duan, W.-X., & An, Q. (2016). Raoultella sp. Strain L03 fixes N2 in association with micropropagated sugarcane plants. Journal of Basic Microbiology, 56(8), 934–940. https://doi.org/10.1002/jobm.201500738

Lupatini, M., Korthals, G. W., de Hollander, M., Janssens, T. K. S., & Kuramae, E. E. (2017). Soil Microbiome Is More Heterogeneous in Organic Than in Conventional Farming System. Frontiers in Microbiology, 7. https://doi.org/10.3389/fmicb.2016.02064

Lutz, K. A., Wang, W., Zdepski, A., & Michael, T. P. (2011). Isolation and analysis of high quality nuclear DNA with reduced organellar DNA for plant genome sequencing and resequencing. BMC Biotechnology, 11(1), 54. https://doi.org/10.1186/1472-6750-11-54

Ma, L., Cao, Y. H., Cheng, M. H., Huang, Y., Mo, M. H., Wang, Y., Yang, J. Z., & Yang, F. X. (2013). Phylogenetic diversity of bacterial endophytes of Panax notoginseng with antagonistic characteristics towards pathogens of root-rot disease complex. Antonie van Leeuwenhoek, 103(2), 299–312. https://doi.org/10.1007/s10482-012-9810-3

Ma, Y., Rajkumar, M., Luo, Y., & Freitas, H. (2011). Inoculation of endophytic bacteria on host and non-host plants—Effects on plant growth and Ni uptake. Journal of Hazardous Materials, 195, 230–237. https://doi.org/10.1016/j.jhazmat.2011.08.034

Ma, Y., Rajkumar, M., Zhang, C., & Freitas, H. (2016). Inoculation of Brassica oxyrrhina with plant growth promoting bacteria for the improvement of heavy metal phytoremediation under drought conditions. Journal of Hazardous Materials, 320, 36–44. https://doi.org/10.1016/j.jhazmat.2016.08.009

Macedo-Raygoza, G. M., Valdez-Salas, B., Prado, F. M., Prieto, K. R., Yamaguchi, L. F., Kato, M. J., Canto-Canché, B. B., Carrillo-Beltrán, M., Di Mascio, P., White, J. F., & Beltrán-García, M. J. (2019). Enterobacter cloacae, an Endophyte That Establishes a Nutrient-Transfer Symbiosis With Banana Plants and Protects Against the Black Sigatoka Pathogen. Frontiers in Microbiology, 10. https://doi.org/10.3389/fmicb.2019.00804

209

Madhaiyan, M., Poonguzhali, S., Lee, J.-S., Lee, K.-C., Saravanan, V. S., & Santhanakrishnan, P. (2010). Microbacterium azadirachtae sp. Nov., a plant- growth-promoting actinobacterium isolated from the rhizoplane of neem seedlings. International Journal of Systematic and Evolutionary Microbiology, 60(7), 1687–1692. https://doi.org/10.1099/ijs.0.015800-0

Magnani, G. S., Didonet, C. M., Cruz, L. M., Picheth, C. F., Pedrosa, F. O., & Souza, E. M. (2010). Diversity of endophytic bacteria in Brazilian sugarcane. Genetics and Molecular Research, 9(1), 250–258. https://doi.org/10.4238/vol9-1gmr703

Mahmood, T., Mehnaz, S., Fleischmann, F., Ali, R., Hashmi, Z. H., & Iqbal, Z. (2014). Soil sterilization effects on root growth and formation of rhizosheaths in wheat seedlings. Pedobiologia, 57(3), 123–130. https://doi.org/10.1016/j.pedobi.2013.12.005

Markova, Yu. A., Romanenko, A. S., & Dukhanina, A. V. (2005). Isolation of Bacteria of the Family Enterobacteriaceae from Plant Tissues. Microbiology, 74(5), 575– 578. https://doi.org/10.1007/s11021-005-0105-9

Marquez-Santacruz, H. A., Hernandez-Leon, R., Orozco-Mosqueda, M. C., Velazquez- Sepulveda, I., & Santoyo, G. (2010). Diversity of bacterial endophytes in roots of Mexican husk tomato plants (Physalis ixocarpa) and their detection in the rhizosphere. Genetics and Molecular Research, 9(4), 2372–2380. https://doi.org/10.4238/vol9-4gmr921

Martin, M. (2011). Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet.Journal, 17(1), 10–12. https://doi.org/10.14806/ej.17.1.200

Maru, A., Haruna, O. A., & Charles Primus, W. (2015). Coapplication of Chicken Litter Biochar and Urea Only to Improve Nutrients Use Efficiency and Yield of Oryza sativa L. Cultivation on a Tropical Acid Soil [Research Article]. The Scientific World Journal; Hindawi. https://doi.org/10.1155/2015/943853

Mašková, T., & Herben, T. (2018). Root:shoot ratio in developing seedlings: How seedlings change their allocation in response to seed mass and ambient nutrient supply. Ecology and Evolution, 8(14), 7143–7150. https://doi.org/10.1002/ece3.4238

Massart, S., Martinez-Medina, M., & Jijakli, M. H. (2015). Biological control in the microbiome era: Challenges and opportunities. Biological Control, 89, 98–108. https://doi.org/10.1016/j.biocontrol.2015.06.003 Mayak, S., Tirosh, T., & Glick, B. R. (2001). Stimulation of the Growth of Tomato,

210

Pepper and Mung Bean Plants by the Plant Growth- Promoting Bacterium Enterobacter cloacae CAL3. Biological Agriculture & Horticulture, 19(3), 261–274. https://doi.org/10.1080/01448765.2001.9754929

McCarty, G. W., Lyssenko, N. N., & Starr, J. L. (1998). Short-term Changes in Soil Carbon and Nitrogen Pools during Tillage Management Transition. Soil Science Society of America Journal, 62(6), 1564–1571. https://doi.org/10.2136/sssaj1998.03615995006200060013x

McGilloway, D. A. (2005). Evolution of integrated crop-livestock production systems. In Grassland: A global resource: A Global Resource (pp. 137–148). Wageningen Academic Publishers.

McKnight, D. T., Huerlimann, R., Bower, D. S., Schwarzkopf, L., Alford, R. A., & Zenger, K. R. (2019). Methods for normalizing microbiome data: An ecological perspective. Methods in Ecology and Evolution, 10(3), 389–400. https://doi.org/10.1111/2041-210X.13115

McMurdie, P. J., & Holmes, S. (2013). phyloseq: An R Package for Reproducible Interactive Analysis and Graphics of Microbiome Census Data. PLOS ONE, 8(4), e61217. https://doi.org/10.1371/journal.pone.0061217

McMurdie, P. J., & Holmes, S. (2014). Waste Not, Want Not: Why Rarefying Microbiome Data Is Inadmissible. PLOS Computational Biology, 10(4), e1003531. https://doi.org/10.1371/journal.pcbi.1003531

McNeil, M. M., Brown, J. M., Carvalho, M. E., Hollis, D. G., Morey, R. E., & Reller, L. B. (2004). Molecular epidemiologic evaluation of endocarditis due to Oerskovia turbata and CDC group A-3 associated with contaminated homograft valves. Journal of Clinical Microbiology, 42(6), 2495–2500. https://doi.org/10.1128/JCM.42.6.2495-2500.2004

Melnick, R. L., Zidack, N. K., Bailey, B. A., Maximova, S. N., Guiltinan, M., & Backman, P. A. (2008). Bacterial endophytes: Bacillus spp. from annual crops as potential biological control agents of black pod rot of cacao. Biological Control, 46(1), 46–56. https://doi.org/10.1016/j.biocontrol.2008.01.022

Michalk, D. L., Badgery, W. B., & Kemp, D. R. (2017). Balancing animal, pasture and environmental outcomes in grazing management experiments. Animal Production Science, 57(9), 1775. https://doi.org/10.1071/AN16132

Miliute, I., Buzaite, O., Baniulis, D., & Stanys, V. (2015). Bacterial endophytes in agricultural crops and their role in stress tolerance: A review. Zemdirbyste- Agriculture, 102, 465–478. https://doi.org/10.13080/z-a.2015.102.060 211

Miransari, M., Bahrami, H. A., Rejali, F., & Malakouti, M. J. (2009). Effects of arbuscular mycorrhiza, soil sterilization, and soil compaction on wheat (Triticum aestivum L.) nutrients uptake. Soil and Tillage Research, 104(1), 48– 55. https://doi.org/10.1016/j.still.2008.11.006

Mishra, M., Patole, S., & Mohapatra, H. (2017). Draft Genome Sequences of Nonclinical and Clinical Enterobacter cloacae Isolates Exhibiting Multiple Antibiotic Resistance and Virulence Factors. Genome Announcements, 5(45). https://doi.org/10.1128/genomeA.01218-17

Moe, K., Mg, K. W., Win, K. K., & Yamakawa, T. (2017). Effects of Combined Application of Inorganic Fertilizer and Organic Manures on Nitrogen Use and Recovery Efficiencies of Hybrid Rice (Palethwe-1). American Journal of Plant Sciences, 08(05), 1043. https://doi.org/10.4236/ajps.2017.85069

Mollet, C., Drancourt, M., & Raoult, D. (1997). RpoB sequence analysis as a novel basis for bacterial identification. Molecular Microbiology, 26(5), 1005–1011. https://doi.org/10.1046/j.1365-2958.1997.6382009.x

Montero, E., Cabot, C., Poschenrieder, C., & Barcelo, J. (1998). Relative importance of osmotic-stress and ion-specific effects on ABA-mediated inhibition of leaf expansion growth in Phaseolus vulgaris. Plant, Cell & Environment, 21(1), 54– 62. https://doi.org/10.1046/j.1365-3040.1998.00249.x

Morella, N. M., Weng, F. C.-H., Joubert, P. M., Metcalf, C. J. E., Lindow, S., & Koskella, B. (2020). Successive passaging of a plant-associated microbiome reveals robust habitat and host genotype-dependent selection. Proceedings of the National Academy of Sciences, 117(2), 1148–1159. https://doi.org/10.1073/pnas.1908600116

Morella, N. M., Zhang, X., & Koskella, B. (2019). Tomato Seed-Associated Bacteria Confer Protection of Seedlings Against Foliar Disease Caused by Pseudomonas syringae. Phytobiomes Journal, 3(3), 177–190. https://doi.org/10.1094/PBIOMES-01-19-0007-R

Nandhini, S., Sendhilvel, V., & Babu, S. (2012). Endophytic bacteria from tomato and their efficacy against Fusarium oxysporum f. Sp. Lycopersici, the wilt pathogen. Journal of Biopesticides, 5(2), 178.

Nawangsih, A. A., Damayanti, I., Wiyono, S., & Kartika, J. G. (2011). Selection and Characterization of Endophytic Bacteria as Biocontrol Agents of Tomato Bacterial Wilt Disease. HAYATI Journal of Biosciences, 18(2), 66–70. https://doi.org/10.4308/hjb.18.2.66

212

Needham, P. (1973). Nutritional disorders. The UK Tomato Manual. Grower Books, London.

Niu, B., Paulson, J. N., Zheng, X., & Kolter, R. (2017). Simplified and representative bacterial community of maize roots. Proceedings of the National Academy of Sciences, 114(12), E2450–E2459. https://doi.org/10.1073/pnas.1616148114

Ogier, J.-C., Pages, S., Galan, M., Barret, M., & Gaudriault, S. (2019). RpoB, a promising marker for analyzing the diversity of bacterial communities by amplicon sequencing. BioRxiv, 626119. https://doi.org/10.1101/626119

Oksanen, J., Blanchet, F. G., Friendly, M., Kindt, R., Legendre, P., McGlinn, D., Minchin, P. R., O’Hara, R. B., Simpson, G. L., Solymos, P., Stevens, M. H. H., Szoecs, E., & Wagner, H. (2019). vegan: Community Ecology Package (Version R package version 2.5-6) [Computer software]. https://CRAN.R- project.org/package=vegan

Omeira, N., Barbour, E. K., Nehme, P. A., Hamadeh, S. K., Zurayk, R., & Bashour, I. (2006). Microbiological and chemical properties of litter from different chicken types and production systems. Science of The Total Environment, 367(1), 156– 162. https://doi.org/10.1016/j.scitotenv.2006.02.019

Ottesen, A. R., González Peña, A., White, J. R., Pettengill, J. B., Li, C., Allard, S., Rideout, S., Allard, M., Hill, T., Evans, P., Strain, E., Musser, S., Knight, R., & Brown, E. (2013). Baseline survey of the anatomical microbial ecology of an important food plant: Solanum lycopersicum (tomato). BMC Microbiology, 13(1), 114. https://doi.org/10.1186/1471-2180-13-114

Overmann, J., Abt, B., & Sikorski, J. (2017). Present and Future of Culturing Bacteria. Annual Review of Microbiology, 71(1), 711–730. https://doi.org/10.1146/annurev-micro-090816-093449

Özdemir, F., & Arslan, S. (2019). Molecular Characterization and Toxin Profiles of Bacillus spp. Isolated from Retail Fish and Ground Beef. Journal of Food Science, 84(3), 548–556. https://doi.org/10.1111/1750-3841.14445

Palleroni, N. J. (1997). Prokaryotic diversity and the importance of culturing. Antonie Van Leeuwenhoek, 72(1), 3–19. https://doi.org/10.1023/a:1000394109961

Palmieri, D., Vitullo, D., De Curtis, F., & Lima, G. (2017). A microbial consortium in the rhizosphere as a new biocontrol approach against fusarium decline of chickpea. Plant and Soil, 412(1–2), 425–439. https://doi.org/10.1007/s11104- 016-3080-1

213

Pan, X., Richardson, M. D., Deng, S., Kremer, R. J., English, J. T., Mihail, J. D., Sams, C. E., Scharf, P. C., Veum, K. S., & Xiong, X. (2017). Effect of Organic Amendment and Cultural Practice on Large Patch Occurrence and Soil Microbial Community. Crop Science, 57(4), 2263–2272. https://doi.org/10.2135/cropsci2016.09.0809

Panke-Buisse, K., Poole, A. C., Goodrich, J. K., Ley, R. E., & Kao-Kniffin, J. (2015). Selection on soil microbiomes reveals reproducible impacts on plant function. The ISME Journal, 9(4), 980–989. https://doi.org/10.1038/ismej.2014.196

Park, H. Y., Kim, K. K., Jin, L., & Lee, S.-T. (2006). Microbacterium paludicola sp. Nov., a novel xylanolytic bacterium isolated from swamp forest. International Journal of Systematic and Evolutionary Microbiology, 56(Pt 3), 535–539. https://doi.org/10.1099/ijs.0.63945-0

Paulson, J. N., Stine, O. C., Bravo, H. C., & Pop, M. (2013). Differential abundance analysis for microbial marker-gene surveys. Nature Methods, 10(12), 1200– 1202. https://doi.org/10.1038/nmeth.2658

Pavlova, A. S., Leontieva, M. R., Smirnova, T. A., Kolomeitseva, G. L., Netrusov, A. I., & Tsavkelova, E. A. (2017). Colonization strategy of the endophytic plant growth-promoting strains of Pseudomonas fluorescens and Klebsiella oxytoca on the seeds, seedlings and roots of the epiphytic orchid, Dendrobium nobile Lindl. Journal of Applied Microbiology, 123(1), 217–232. https://doi.org/10.1111/jam.13481

Pavoine, S., Dufour, A.-B., & Chessel, D. (2004). From dissimilarities among species to dissimilarities among communities: A double principal coordinate analysis. Journal of Theoretical Biology, 228(4), 523–537. https://doi.org/10.1016/j.jtbi.2004.02.014

Peralta, A. L., Sun, Y., McDaniel, M. D., & Lennon, J. T. (2018). Crop rotational diversity increases disease suppressive capacity of soil microbiomes. Ecosphere, 9(5), e02235. https://doi.org/10.1002/ecs2.2235

Postma-Blaauw, M. B., Goede, R. G. M. de, Bloem, J., Faber, J. H., & Brussaard, L. (2010). Soil biota community structure and abundance under agricultural intensification and extensification. Ecology, 91(2), 460–473. https://doi.org/10.1890/09-0666.1

Qingxiang Yang, Hao Zhang, Yuhui Guo, & Tiantian Tian. (2016). Influence of Chicken Manure Fertilization on Antibiotic-Resistant Bacteria in Soil and the Endophytic Bacteria of Pakchoi. International Journal of Environmental Research & Public Health, 13(7), 662. https://doi.org/10.3390/ijerph13070662 214

Quambusch, M., & Winkelmann, T. (2018). Bacterial Endophytes in Plant Tissue Culture: Mode of Action, Detection, and Control. In V. M. Loyola-Vargas & N. Ochoa-Alejo (Eds.), Plant Cell Culture Protocols (pp. 69–88). Springer. https://doi.org/10.1007/978-1-4939-8594-4_4

Quast, C., Pruesse, E., Yilmaz, P., Gerken, J., Schweer, T., Yarza, P., Peplies, J., & Glöckner, F. O. (2013). The SILVA ribosomal RNA gene database project: Improved data processing and web-based tools. Nucleic Acids Research, 41(D1), D590–D596. https://doi.org/10.1093/nar/gks1219

R Core Team. (2014). R: A language and environment for statistical computing. R Foundation for Statistical Computing. http://www.R-project.org/

Radford, B. J., Yule, D. F., McGarry, D., & Playford, C. (2001). Crop responses to applied soil compaction and to compaction repair treatments. Soil and Tillage Research, 61(3), 157–166. https://doi.org/10.1016/S0167-1987(01)00194-5

Ramasamy, D., Mishra, A. K., Lagier, J.-C., Padhmanabhan, R., Rossi, M., Sentausa, E., Raoult, D., & Fournier, P.-E. (2014). A polyphasic strategy incorporating genomic data for the taxonomic description of novel bacterial species. International Journal of Systematic and Evolutionary Microbiology, 64(2), 384– 391. https://doi.org/10.1099/ijs.0.057091-0

Rasche, F., Velvis, H., Zachow, C., Berg, G., Elsas, J. D. V., & Sessitsch, A. (2006). Impact of transgenic potatoes expressing anti‐bacterial agents on bacterial endophytes is comparable with the effects of plant genotype, soil type and pathogen infection. Journal of Applied Ecology, 43(3), 555–566. https://doi.org/10.1111/j.1365-2664.2006.01169.x

Ravindran, B., Mupambwa, H. A., Silwana, S., & Mnkeni, P. N. S. (2017). Assessment of nutrient quality, heavy metals and phytotoxic properties of chicken manure on selected commercial vegetable crops. Heliyon, 3(12), e00493. https://doi.org/10.1016/j.heliyon.2017.e00493

Raymond, B., Johnston, P. R., Nielsen-LeRoux, C., Lereclus, D., & Crickmore, N. (2010). Bacillus thuringiensis: An impotent pathogen? Trends in Microbiology, 18(5), 189–194. https://doi.org/10.1016/j.tim.2010.02.006

Reasoner, D. J., & Geldreich, E. E. (1985). A new medium for the enumeration and subculture of bacteria from potable water. Applied and Environmental Microbiology, 49(1), 1–7.

215

Ren, G., Zhang, H., Lin, X., Zhu, J., & Jia, Z. (2015). Response of leaf endophytic bacterial community to elevated CO2 at different growth stages of rice plant. Frontiers in Microbiology, 6. https://doi.org/10.3389/fmicb.2015.00855

Riegel, C., & Noe, J. P. (2000). Chicken Litter Soil Amendment Effects on Soilborne Microbes and Meloidogyne incognita on Cotton. Plant Disease, 84(12), 1275– 1281. https://doi.org/10.1094/PDIS.2000.84.12.1275

Romero, F. M., Marina, M., & Pieckenstain, F. L. (2014). The communities of tomato (Solanum lycopersicum L.) leaf endophytic bacteria, analyzed by 16S-ribosomal RNA gene pyrosequencing. FEMS Microbiology Letters, 351(2), 187–194. https://doi.org/10.1111/1574-6968.12377

Rosenblueth, M., & Martínez-Romero, E. (2006). Bacterial Endophytes and Their Interactions with Hosts. Molecular Plant-Microbe Interactions®, 19(8), 827– 837. https://doi.org/10.1094/MPMI-19-0827

Rosenzweig, N., Tiedje, J. M., Quensen, J. F., Meng, Q., & Hao, J. J. (2011). Microbial Communities Associated with Potato Common Scab-Suppressive Soil Determined by Pyrosequencing Analyses. Plant Disease, 96(5), 718–725. https://doi.org/10.1094/PDIS-07-11-0571

Roy, S., & Kashem, M. A. (2014). Effects of Organic Manures in Changes of Some Soil Properties at Different Incubation Periods. Open Journal of Soil Science, 2014. https://doi.org/10.4236/ojss.2014.43011

Russelle, M. P., Entz, M. H., & Franzluebbers, A. J. (2007). Reconsidering integrated crop–livestock systems in North America. Agronomy Journal, 99(2), 325–334.

Sahoo, U. K., Singh, S. L., Gogoi, A., Kenye, A., & Sahoo, S. S. (2019). Active and passive soil organic carbon pools as affected by different land use types in Mizoram, Northeast India. PLoS ONE, 14(7). https://doi.org/10.1371/journal.pone.0219969

Sainju, U. M., Dris, R., & Singh, B. (2003). Mineral nutrition of tomato. 9.

Saleem, M., Law, A. D., Sahib, M. R., Pervaiz, Z. H., & Zhang, Q. (2018). Impact of root system architecture on rhizosphere and root microbiome. Rhizosphere, 6, 47–51. https://doi.org/10.1016/j.rhisph.2018.02.003

Sanders, E. R., & Miller, J. H. (2010). I, Microbiologist: A Discovery-Based Course In Microbial Ecology and Molecular Evolution. ASM Press.

216

Santos, S. R., & Ochman, H. (2004). Identification and phylogenetic sorting of bacterial lineages with universally conserved genes and proteins. Environmental Microbiology, 6(7), 754–759. https://doi.org/10.1111/j.1462-2920.2004.00617.x

Santoyo, G., Moreno-Hagelsieb, G., del Carmen Orozco-Mosqueda, Ma., & Glick, B. R. (2016). Plant growth-promoting bacterial endophytes. Microbiological Research, 183, 92–99. https://doi.org/10.1016/j.micres.2015.11.008

Schiere, H., & Kater, L. (2001). Mixed crop-livestock farming. A review of traditional technologies based on literature and field experience. FAO Animal Production and Health Paper (FAO). https://agris.fao.org/agris- search/search.do?recordID=XF2003410307

Schliep, K., Potts, A. J., Morrison, D. A., & Grimm, G. W. (2017). Intertwining phylogenetic trees and networks. Methods in Ecology and Evolution, 8(10), 1212–1220. https://doi.org/10.1111/2041-210X.12760

Schneider, C. A., Rasband, W. S., & Eliceiri, K. W. (2012). NIH Image to ImageJ: 25 years of image analysis. Nature Methods, 9(7), 671–675. https://doi.org/10.1038/nmeth.2089

Serrasolsas, I., & Khanna, P. K. (1995). Changes in heated and autoclaved forest soils of S.E. Australia. II. Phosphorus and phosphatase activity. Biogeochemistry, 29(1), 25–41. https://doi.org/10.1007/BF00002592

Sexton, W. K., Fidero, M., Spain, J. C., Jiang, L., Bucalo, K., Cruse-Sanders, J. M., & Pullman, G. S. (2020). Characterization of endophytic bacterial communities within greenhouse and field-grown rhizomes of three rare pitcher plant species (Sarracenia oreophila, S. leucophylla, and S. purpurea spp. Venosa) with an emphasis on nitrogen-fixing bacteria. Plant and Soil, 447(1), 257–279. https://doi.org/10.1007/s11104-019-04372-8

Shahzad, R., Khan, A. L., Bilal, S., Asaf, S., & Lee, I.-J. (2017). Plant growth- promoting endophytic bacteria versus pathogenic infections: An example of Bacillus amyloliquefaciens RWL-1 and Fusarium oxysporum f. sp. lycopersici in tomato. PeerJ, 5, e3107. https://doi.org/10.7717/peerj.3107

Shahzad, R., Khan, A. L., Bilal, S., Waqas, M., Kang, S.-M., & Lee, I.-J. (2017). Inoculation of abscisic acid-producing endophytic bacteria enhances salinity stress tolerance in Oryza sativa. Environmental and Experimental Botany, 136, 68–77. https://doi.org/10.1016/j.envexpbot.2017.01.010

217

Shaik, S. P., & Thomas, P. (2019). In Vitro Activation of Seed-Transmitted Cultivation- Recalcitrant Endophytic Bacteria in Tomato and Host–Endophyte Mutualism. Microorganisms, 7(5). https://doi.org/10.3390/microorganisms7050132

Shannon, C. E. (1948). A Mathematical Theory of Communication. Bell System Technical Journal, 27(4), 623–656. https://doi.org/10.1002/j.1538- 7305.1948.tb00917.x

Sharon, J. A., Hathwaik, L. T., Glenn, G. M., Imam, S. H., & Lee, C. C. (2016). Isolation of efficient phosphate solubilizing bacteria capable of enhancing tomato plant growth. Journal of Soil Science and Plant Nutrition, 16(2), 525– 536. https://doi.org/10.4067/S0718-95162016005000043

Sharp, R. E., & LeNoble, M. E. (2002). ABA, ethylene and the control of shoot and root growth under water stress. Journal of Experimental Botany, 53(366), 33–37. https://doi.org/10.1093/jexbot/53.366.33

Shi, Y., Lou, K., & Li, C. (2010). Growth and photosynthetic efficiency promotion of sugar beet (Beta vulgaris L.) by endophytic bacteria. Photosynthesis Research, 105(1), 5–13. https://doi.org/10.1007/s11120-010-9547-7

Simon, H. M., Smith, K. P., Dodsworth, J. A., Guenthner, B., Handelsman, J., & Goodman, R. M. (2001). Influence of Tomato Genotype on Growth of Inoculated and Indigenous Bacteria in the Spermosphere. Applied and Environmental Microbiology, 67(2), 514–520. https://doi.org/10.1128/AEM.67.2.514-520.2001

Sistani, K. R., Torbert, H. A., Way, T. R., Bolster, C. H., Pote, D. H., & Warren, J. G. (2009). Broiler Litter Application Method and Runoff Timing Effects on Nutrient and Escherichia coli Losses from Tall Fescue Pasture. Journal of Environmental Quality, 38(3), 1216–1223. https://doi.org/10.2134/jeq2008.0185

Sogin, M. L., Morrison, H. G., Huber, J. A., Mark Welch, D., Huse, S. M., Neal, P. R., Arrieta, J. M., & Herndl, G. J. (2006). Microbial diversity in the deep sea and the underexplored “rare biosphere.” Proceedings of the National Academy of Sciences of the United States of America, 103(32), 12115–12120. https://doi.org/10.1073/pnas.0605127103

Soil Fertility Lab, O. S. U. (2019). Procedures for the Determination of Permanganate Oxidizable Carbon, Autoclaved-Citrate Extractable (ACE) Soil Protein and Soil Respiration. https://soilfertility.osu.edu/protocols

Soil Survey Staff. (2019). Natural Resources Conservation Service, United States Department of Agriculture. http://websoilsurvey.sc.egov.usda.gov/ 218

Souza, S. A., Xavier, A. A., Costa, M. R., Cardoso, A. M. S., Pereira, M. C. T., & Nietsche, S. (2013). Endophytic bacterial diversity in banana ‘Prata Anã’ (Musa spp.) roots. Genetics and Molecular Biology, 36(2), 252–264. https://doi.org/10.1590/S1415-47572013000200016

Sulc, R. M., & Franzluebbers, A. J. (2014). Exploring integrated crop–livestock systems in different ecoregions of the United States. European Journal of Agronomy, 57, 21–30. https://doi.org/10.1016/j.eja.2013.10.007

Sun, H. Y., Deng, S. P., & Raun, W. R. (2004). Bacterial Community Structure and Diversity in a Century-Old Manure-Treated Agroecosystem. Applied and Environmental Microbiology, 70(10), 5868–5874. https://doi.org/10.1128/AEM.70.10.5868-5874.2004

Szymańska, S., Borruso, L., Brusetti, L., Hulisz, P., Furtado, B., & Hrynkiewicz, K. (2018). Bacterial microbiome of root-associated endophytes of Salicornia europaea in correspondence to different levels of salinity. Environmental Science and Pollution Research, 25(25), 25420–25431. https://doi.org/10.1007/s11356-018-2530-0

Tamošiūnė, I., Stanienė, G., Haimi, P., Stanys, V., Rugienius, R., & Baniulis, D. (2018). Endophytic Bacillus and Pseudomonas spp. Modulate Apple Shoot Growth, Cellular Redox Balance, and Protein Expression Under in Vitro Conditions. Frontiers in Plant Science, 9. https://doi.org/10.3389/fpls.2018.00889

Tamura, K., & Nei, M. (1993). Estimation of the number of nucleotide substitutions in the control region of mitochondrial DNA in humans and chimpanzees. Molecular Biology and Evolution, 10(3), 512–526. https://doi.org/10.1093/oxfordjournals.molbev.a040023

Tamura, K., Nei, M., & Kumar, S. (2004). Prospects for inferring very large phylogenies by using the neighbor-joining method. Proceedings of the National Academy of Sciences of the United States of America, 101(30), 11030–11035. https://doi.org/10.1073/pnas.0404206101

Tariq, M., Hameed, S., Yasmeen, T., Zahid, M., & Zafar, M. (2014). Molecular characterization and identification of plant growth promoting endophytic bacteria isolated from the root nodules of pea (Pisum sativum L.). World Journal of Microbiology and Biotechnology, 30(2), 719–725. https://doi.org/10.1007/s11274-013-1488-9

Telias, A., White, J. R., Pahl, D. M., Ottesen, A. R., & Walsh, C. S. (2011). Bacterial community diversity and variation in spray water sources and the tomato fruit surface. BMC Microbiology, 11, 81. https://doi.org/10.1186/1471-2180-11-81 219

Thokchom, E., Thakuria, D., Kalita, M. C., Sharma, C. K., & Talukdar, N. C. (2017). Root colonization by host-specific rhizobacteria alters indigenous root endophyte and rhizosphere soil bacterial communities and promotes the growth of mandarin orange. European Journal of Soil Biology, 79, 48–56. https://doi.org/10.1016/j.ejsobi.2017.02.003

Thomas, P., & Upreti, R. (2016). Evaluation of tomato seedling root-associated bacterial endophytes towards organic seedling production. Organic Agriculture, 6(2), 89–98. https://doi.org/10.1007/s13165-015-0111-9

Tian, B.-Y., Cao, Y., & Zhang, K.-Q. (2015). Metagenomic insights into communities, functions of endophytes and their associates with infection by root-knot nematode, Meloidogyne incognita, in tomato roots. Scientific Reports, 5(1). https://doi.org/10.1038/srep17087

Tian, B., Zhang, C., Ye, Y., Wen, J., Wu, Y., Wang, H., Li, H., Cai, S., Cai, W., Cheng, Z., Lei, S., Ma, R., Lu, C., Cao, Y., Xu, X., & Zhang, K. (2017). Beneficial traits of bacterial endophytes belonging to the core communities of the tomato root microbiome. Agriculture Ecosystems & Environment, 247, 149–156. https://doi.org/10.1016/j.agee.2017.06.041

Toju, H., Okayasu, K., & Notaguchi, M. (2019). Leaf-associated microbiomes of grafted tomato plants. Scientific Reports, 9(1), 1787. https://doi.org/10.1038/s41598-018-38344-2

Tracy, S. R., Black, C. R., Roberts, J. A., Sturrock, C., Mairhofer, S., Craigon, J., & Mooney, S. J. (2012). Quantifying the impact of soil compaction on root system architecture in tomato (Solanum lycopersicum) by X-ray micro-computed tomography. Annals of Botany, 110(2), 511–519. https://doi.org/10.1093/aob/mcs031

Udo, H. M. J., Aklilu, H. A., Phong, L. T., Bosma, R. H., Budisatria, I. G. S., Patil, B. R., Samdup, T., & Bebe, B. O. (2011). Impact of intensification of different types of livestock production in smallholder crop-livestock systems. Livestock Science, 139(1), 22–29. https://doi.org/10.1016/j.livsci.2011.03.020

USDA-AMS. (2019). USDA: Agricultural Marketing Service. Agricultural Marketing Service. https://www.ams.usda.gov/

Vos, M., Quince, C., Pijl, A. S., de Hollander, M., & Kowalchuk, G. A. (2012). A comparison of rpoB and 16S rRNA as markers in pyrosequencing studies of bacterial diversity. PloS One, 7(2), e30600. https://doi.org/10.1371/journal.pone.0030600

220

Wagner, M. R., Lundberg, D. S., Del Rio, T. G., Tringe, S. G., Dangl, J. L., & Mitchell- Olds, T. (2016). Host genotype and age shape the leaf and root microbiomes of a wild perennial plant. Nature Communications, 7, 12151. https://doi.org/10.1038/ncomms12151

Walker, J. J., & Pace, N. R. (2007). Phylogenetic Composition of Rocky Mountain Endolithic Microbial Ecosystems. Applied and Environmental Microbiology, 73(11), 3497–3504. https://doi.org/10.1128/AEM.02656-06

Walpola, B., & Yoon, M.-H. (2013). Isolation and characterization of phosphate solubilizing bacteria and their co-inoculation efficiency on tomato plant growth and phosphorous uptake. Afr J Microbiol Res, 7, 266–275. https://doi.org/10.5897/AJMR12.2282

Walterson, A. M., & Stavrinides, J. (2015). Pantoea: Insights into a highly versatile and diverse genus within the Enterobacteriaceae. FEMS Microbiology Reviews, 39(6), 968–984. https://doi.org/10.1093/femsre/fuv027

Wander, M. (2004). Soil Organic Matter Fractions and Their Relevance to Soil Function. In F. Magdoff & R. Weil (Eds.), Soil Organic Matter in Sustainable Agriculture (Vol. 20042043). CRC Press. https://doi.org/10.1201/9780203496374.ch3

Wander, M. M., Traina, S. J., Stinner, B. R., & Peters, S. E. (1994). Organic and Conventional Management Effects on Biologically Active Soil Organic Matter Pools. Soil Science Society of America Journal, 58(4), 1130–1139. https://doi.org/10.2136/sssaj1994.03615995005800040018x

Wang, E. T., Tan, Z. Y., Guo, X. W., Rodríguez-Duran, R., Boll, G., & Martínez- Romero, E. (2006). Diverse endophytic bacteria isolated from a leguminous tree Conzattia multiflora grown in Mexico. Archives of Microbiology, 186(4), 251– 259. https://doi.org/10.1007/s00203-006-0141-5

Wang, Q., Garrity, G. M., Tiedje, J. M., & Cole, J. R. (2007). Naïve Bayesian Classifier for Rapid Assignment of rRNA Sequences into the New Bacterial Taxonomy. Applied and Environmental Microbiology, 73(16), 5261–5267. https://doi.org/10.1128/AEM.00062-07

Wang, S.-S., Liu, J.-M., Sun, J., Sun, Y.-F., Liu, J.-N., Jia, N., Fan, B., & Dai, X.-F. (2019). Diversity of culture-independent bacteria and antimicrobial activity of culturable endophytic bacteria isolated from different Dendrobium stems. Scientific Reports, 9. https://doi.org/10.1038/s41598-019-46863-9

221

Weisburg, W. G., Barns, S. M., Pelletier, D. A., & Lane, D. J. (1991). 16S ribosomal DNA amplification for phylogenetic study. Journal of Bacteriology, 173(2), 697–703. https://doi.org/10.1128/jb.173.2.697-703.1991

Wicaksono, W. A., Jones, E. E., Casonato, S., Monk, J., & Ridgway, H. J. (2018). Biological control of Pseudomonas syringae pv. Actinidiae (Psa), the causal agent of bacterial canker of kiwifruit, using endophytic bacteria recovered from a medicinal plant. Biological Control, 116, 103–112. https://doi.org/10.1016/j.biocontrol.2017.03.003

Wickham, H. (2016). ggplot2: Elegant Graphics for Data Analysis. Springer-Verlag.

Wickham, H., Hester, J., & Francois, R. (2018). readr: Read Rectangular Text Data. (R package version 1.3.1.) [Computer software]. https://CRAN.R- project.org/package=readr

Willis, A., & Bunge, J. (2015). Estimating diversity via frequency ratios. Biometrics, 71(4), 1042–1049. https://doi.org/10.1111/biom.12332

Wolf, D. C., Dao, T. H., Scott, H. D., & Lavy, T. L. (1989). Influence of Sterilization Methods on Selected Soil Microbiological, Physical, and Chemical Properties. Journal of Environmental Quality, 18(1), 39–44. https://doi.org/10.2134/jeq1989.00472425001800010007x

Woods, L. E. (1989). Active organic matter distribution in the surface 15 cm of undisturbed and cultivated soil. Biology and Fertility of Soils, 8(3). https://doi.org/10.1007/BF00266490

Wright, E., S. (2016). Using DECIPHER v2.0 to Analyze Big Biological Sequence Data in R. The R Journal, 8(1), 352. https://doi.org/10.32614/RJ-2016-025

Xia, Y., DeBolt, S., Dreyer, J., Scott, D., & Williams, M. A. (2015). Characterization of culturable bacterial endophytes and their capacity to promote plant growth from plants grown using organic or conventional practices. Frontiers in Plant Science, 6. https://doi.org/10.3389/fpls.2015.00490

Xu, W., Wang, F., Zhang, M., Ou, T., Wang, R., Strobel, G., Xiang, Z., Zhou, Z., & Xie, J. (2019). Diversity of cultivable endophytic bacteria in mulberry and their potential for antimicrobial and plant growth-promoting activities. Microbiological Research, 229, 126328. https://doi.org/10.1016/j.micres.2019.126328

Yamamoto, S., & Harayama, S. (1995). PCR Amplification and Direct Sequencing of gyrB Genes with Universal Primers and Their Application to the Detection and 222

Taxonomic Analysis of Pseudomonas putida Strains. APPL. ENVIRON. MICROBIOL., 61, 7.

Yang, Q., Ren, S., Niu, T., Guo, Y., Qi, S., Han, X., Liu, D., & Pan, F. (2014). Distribution of antibiotic-resistant bacteria in chicken manure and manure- fertilized vegetables. Environmental Science and Pollution Research International; Heidelberg, 21(2), 1231–1241. http://dx.doi.org.proxy.lib.ohio- state.edu/10.1007/s11356-013-1994-1

Yanti, Y., Warnita, & Reflin. (2019). Involvement of Jasmonic Acid in the Induced Systemic Resistance of Tomato against Ralstonia syzigiisub sp. Indonesiensis by Indigenous Endophyte Bacteria. IOP Conference Series: Earth and Environmental Science, 347, 012024. https://doi.org/10.1088/1755- 1315/347/1/012024

Ye, J., Zhang, R., Nielsen, S., Joseph, S. D., Huang, D., & Thomas, T. (2016). A Combination of Biochar–Mineral Complexes and Compost Improves Soil Bacterial Processes, Soil Quality, and Plant Properties. Frontiers in Microbiology, 7. https://doi.org/10.3389/fmicb.2016.00372

Yoon, J.-H., Kang, S.-J., Schumann, P., & Oh, T.-K. (2007). Cellulosimicrobium terreum sp. Nov., isolated from soil. International Journal of Systematic and Evolutionary Microbiology, 57(Pt 11), 2493–2497. https://doi.org/10.1099/ijs.0.64889-0

Zazou, K., Abdelkrim, T., Chama, Z., Kaldi, A., Tounssi, S., & Bouziane, A. (2016). Isolation and characterization of Microbacterium arthrosphaerae and its effect on tomato culture growth. Der Pharm. Lett., 8(8), 1–8.

Zhang, B.-H., Salam, N., Cheng, J., Li, H.-Q., Yang, J.-Y., Zha, D.-M., Guo, Q.-G., & Li, W.-J. (2017). Microbacterium lacusdiani sp. Nov., a phosphate-solubilizing novel actinobacterium isolated from mucilaginous sheath of Microcystis. The Journal of Antibiotics, 70(2), 147–151. https://doi.org/10.1038/ja.2016.125

Zhang, D., Yan, D., Cheng, H., Fang, W., Huang, B., Wang, X., Wang, X., Yan, Y., Ouyang, C., Li, Y., Wang, Q., & Cao, A. (2020). Effects of multi-year biofumigation on soil bacterial and fungal communities and strawberry yield. Environmental Pollution, 256, 113415. https://doi.org/10.1016/j.envpol.2019.113415

Zhang, L., Li, L., Pan, X., Shi, Z., Feng, X., Gong, B., Li, J., & Wang, L. (2018). Enhanced Growth and Activities of the Dominant Functional Microbiota of Chicken Manure Composts in the Presence of Maize Straw. Frontiers in Microbiology, 9. https://doi.org/10.3389/fmicb.2018.01131 223

Zhang, M., Powell, C. A., Guo, Y., Benyon, L., & Duan, Y. (2013). Characterization of the microbial community structure in Candidatus Liberibacter asiaticus-infected citrus plants treated with antibiotics in the field. BMC Microbiology, 13(1), 112. https://doi.org/10.1186/1471-2180-13-112

Zimmer, W., Hundeshagen, B., & Niederau, E. (1994). Demonstration of the indolepyruvate decarboxylase gene homologue in different auxin-producing species of the Enterobacteriaceae. Canadian Journal of Microbiology, 40(12), 1072–1076. https://doi.org/10.1139/m94-170

224

Appendix A. Supplementary Table

225

Table A.1. Identification of differentially abundant taxa (p<0.1) by comparing chicken grazed pasture plots with pasture only plots.

Sample Origin ASV P Value Phylum Class Order Family Genus (Sampling Month) Rhizosphere Soil (July) 317 2.1x10-6 Actinobacteria Thermoleophilia Solirubrobacterales 67-14 NA 1159 7.5x10-6 Actinobacteria Thermoleophilia Solirubrobacterales 67-14 NA 1371 8.6x10-5 Firmicutes Alicyclobacillaceae Tumebacillus 2130 9.7x10-5 Firmicutes Bacilli Bacillales Alicyclobacillaceae Tumebacillus 1719 1.4x10-4 Actinobacteria Thermoleophilia Solirubrobacterales Solirubrobacteraceae NA 2024 1.4x10-4 Actinobacteria Thermoleophilia Solirubrobacterales 67-14 NA 2325 1.7x10-4 Proteobacteria Betaproteobacteriales TRA3-20 NA 1806 1.8x10-4 Actinobacteria Thermoleophilia Solirubrobacterales 67-14 NA 226 1790 2.3x10-4 Firmicutes Bacilli Bacillales Alicyclobacillaceae Tumebacillus

647 2.8x10-4 Nitrospirae Nitrospira Nitrospirales Nitrospiraceae Nitrospira 954 4.7x10-4 Proteobacteria Gammaproteobacteria Betaproteobacteriales Nitrosomonadaceae mle1-7 1217 4.8x10-4 Proteobacteria Gammaproteobacteria Betaproteobacteriales A21b NA 76 4.9x10-4 Proteobacteria Gammaproteobacteria Xanthomonadales Xanthomonadaceae Stenotrophomonas 758 5.5x10-4 Proteobacteria Alphaproteobacteria Sphingomonadales Sphingomonadaceae Sphingomonas Rhizobiales_Incertae_S

1922 5.6x10-4 Proteobacteria Alphaproteobacteria Rhizobiales edis Alsobacter Allorhizobium-Neorhizobium-

1875 5.6x10-4 Proteobacteria Alphaproteobacteria Rhizobiales Rhizobiaceae Pararhizobium-Rhizobium 2260 6.0x10-4 Chloroflexi Chloroflexia Chloroflexales Roseiflexaceae NA 689 6.1x10-4 Firmicutes Bacilli Bacillales Alicyclobacillaceae Tumebacillus 789 6.3x10-4 Proteobacteria Alphaproteobacteria Rhizobiales Hyphomicrobiaceae Hyphomicrobium 298 6.3x10-4 Proteobacteria Alphaproteobacteria Rhizobiales Xanthobacteraceae Bradyrhizobium Continued 226

Table A.1. Continued

Sample Origin ASV P Value Phylum Class Order Family Genus (Sampling Month) Rhizosphere Soil (July) 2554 6.7x10-4 Actinobacteria Actinobacteria Demequinaceae 1711 7.2x10-4 Proteobacteria Gammaproteobacteria Betaproteobacteriales A21b NA 378 7.9x10-4 Proteobacteria Alphaproteobacteria Rhizobiales Rhodomicrobiaceae Rhodomicrobium 155 8.8x10-4 Actinobacteria Actinobacteria Micrococcales Microbacteriaceae Microbacterium 2807 9.5x10-4 Verrucomicrobia Verrucomicrobiae Pedosphaerales Pedosphaeraceae ADurb.Bin063-1 478 9.6x10-4 Proteobacteria Alphaproteobacteria Dongiales Dongiaceae Dongia 119 9.9x10-4 Firmicutes Bacilli Bacillales Alicyclobacillaceae Tumebacillus 1035 1.1x10-3 Firmicutes Bacilli Bacillales Alicyclobacillaceae Tumebacillus 1065 1.2x10-3 Acidobacteria Subgroup_5 NA NA NA 612 1.3x10-3 Proteobacteria Alphaproteobacteria Micropepsales Micropepsaceae NA 227 656 1.3x10-3 Proteobacteria Alphaproteobacteria Sphingomonadales Sphingomonadaceae Sphingomonas

1105 1.3x10-3 Actinobacteria Actinobacteria Micrococcales Microbacteriaceae Microbacterium 1038 1.4x10-3 Bacteroidetes Bacteroidia Chitinophagales Saprospiraceae NA 1874 1.5x10-3 Actinobacteria Thermoleophilia Gaiellales NA NA 1931 1.6x10-3 Proteobacteria Gammaproteobacteria Gammaproteobacteria_Incertae_Sedis Unknown_Family Acidibacter 2005 1.6x10-3 Bacteroidetes Bacteroidia Chitinophagales Chitinophagaceae NA 205 1.6x10-3 Actinobacteria Actinobacteria Corynebacteriales Mycobacteriaceae Mycobacterium 1299 1.6x10-3 Actinobacteria Actinobacteria Kineosporiales Kineosporiaceae Angustibacter 1943 1.8x10-3 Acidobacteria Acidobacteriia Acidobacteriales NA NA

Continued

227

Table A.1. Continued Sample Origin ASV P Value Phylum Class Order Family Genus (Sampling Month) Rhizosphere Soil (July) 574 1.9x10-3 Actinobacteria Actinobacteria Corynebacteriales Mycobacteriaceae Mycobacterium 2556 2.0x10-3 Proteobacteria Alphaproteobacteria Rhizobiales Xanthobacteraceae Rhodoplanes 1691 2.0x10-3 Acidobacteria Thermoanaerobaculia Thermoanaerobaculales Thermoanaerobaculaceae Subgroup_10 1770 2.2x10-3 Proteobacteria Deltaproteobacteria Myxococcales Blfdi19 NA 2210 2.3x10-3 Actinobacteria Actinobacteria Micrococcales Promicromonosporaceae Cellulosimicrobium 374 2.4x10-3 Firmicutes Bacilli Bacillales Alicyclobacillaceae Tumebacillus 159 2.4x10-3 Firmicutes Bacilli Bacillales Family_XII Exiguobacterium 2346 2.4x10-3 Actinobacteria Thermoleophilia Solirubrobacterales Solirubrobacteraceae Conexibacter 3454 2.4x10-3 Gemmatimonadetes Gemmatimonadetes Gemmatimonadales Gemmatimonadaceae NA 1419 2.4x10-3 Bacteroidetes Bacteroidia Chitinophagales Chitinophagaceae Ferruginibacter 52 2.5x10-3 Proteobacteria Gammaproteobacteria Betaproteobacteriales Burkholderiaceae Pseudoduganella 1426 2.5x10-3 Actinobacteria Thermoleophilia Solirubrobacterales Solirubrobacteraceae Solirubrobacter 228 364 2.5x10-3 Actinobacteria Thermoleophilia Gaiellales Gaiellaceae Gaiella

397 2.6x10-3 Nitrospirae Nitrospira Nitrospirales Nitrospiraceae Nitrospira

Continued

228

Table A.1. Continued Sample Origin ASV P Value Phylum Class Order Family Genus (Sampling Month) Allorhizobium- Stem (July) Neorhizobium- 30 6.9x10-13 Proteobacteria Alphaproteobacteria Rhizobiales Rhizobiaceae Pararhizobium-Rhizobium 701 9.4x10-5 Actinobacteria Actinobacteria Kineosporiales Kineosporiaceae Kineococcus 226 3.3x10-4 Deinococcus-Thermus Deinococci Deinococcales Deinococcaceae Deinococcus 1121 1.3x10-3 Proteobacteria Gammaproteobacteria Betaproteobacteriales Burkholderiaceae Variovorax 468 1.5x10-3 Proteobacteria Alphaproteobacteria Sphingomonadales Sphingomonadaceae Sphingomonas 103 1.8x10-3 Proteobacteria Alphaproteobacteria Rhizobiales Beijerinckiaceae Methylobacterium 140 1.9x10-3 Proteobacteria Alphaproteobacteria Sphingomonadales Sphingomonadaceae Sphingomonas 678 2.3x10-3 Proteobacteria Gammaproteobacteria Betaproteobacteriales Burkholderiaceae Xenophilus -3

229 91 2.5x10 Proteobacteria Gammaproteobacteria Betaproteobacteriales Burkholderiaceae Variovorax 352 3.6x10-3 Actinobacteria Actinobacteria Kineosporiales Kineosporiaceae Kineococcus

469 6.0x10-3 Proteobacteria Alphaproteobacteria Sphingomonadales Sphingomonadaceae Sphingomonas 252 7.9x10-3 Proteobacteria Gammaproteobacteria Betaproteobacteriales Burkholderiaceae Massilia 336 8.0x10-3 Proteobacteria Gammaproteobacteria Betaproteobacteriales Burkholderiaceae Massilia

Root (September) 141 1.0x10-6 Proteobacteria Gammaproteobacteria Betaproteobacteriales Burkholderiaceae Massilia

229