Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

119
Graduate eses and Dissertations Iowa State University Capstones, eses and Dissertations 2011 Biogeography of denitrifying bacteria in a prairie and agro-ecosystem Alescia Ann Roberto Iowa State University Follow this and additional works at: hps://lib.dr.iastate.edu/etd Part of the Ecology and Evolutionary Biology Commons is esis is brought to you for free and open access by the Iowa State University Capstones, eses and Dissertations at Iowa State University Digital Repository. It has been accepted for inclusion in Graduate eses and Dissertations by an authorized administrator of Iowa State University Digital Repository. For more information, please contact [email protected]. Recommended Citation Roberto, Alescia Ann, "Biogeography of denitrifying bacteria in a prairie and agro-ecosystem" (2011). Graduate eses and Dissertations. 12035. hps://lib.dr.iastate.edu/etd/12035

Transcript of Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

Page 1: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

Graduate Theses and Dissertations Iowa State University Capstones, Theses andDissertations

2011

Biogeography of denitrifying bacteria in a prairieand agro-ecosystemAlescia Ann RobertoIowa State University

Follow this and additional works at: https://lib.dr.iastate.edu/etd

Part of the Ecology and Evolutionary Biology Commons

This Thesis is brought to you for free and open access by the Iowa State University Capstones, Theses and Dissertations at Iowa State University DigitalRepository. It has been accepted for inclusion in Graduate Theses and Dissertations by an authorized administrator of Iowa State University DigitalRepository. For more information, please contact [email protected].

Recommended CitationRoberto, Alescia Ann, "Biogeography of denitrifying bacteria in a prairie and agro-ecosystem" (2011). Graduate Theses andDissertations. 12035.https://lib.dr.iastate.edu/etd/12035

Page 2: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

Biogeography of denitrifying bacteria in a

prairie and agro-ecosystem

Alescia Ann Roberto

A thesis submitted to the graduate faculty

in partial fulfillment of the requirements for the degree of

MASTER OF SCIENCE

Major: Microbiology

Program of Study Committee:

Kirsten Hofmockel, Major Professor

Larry Halverson

Thomas Loynachan

Iowa State University

Ames, Iowa

2011

Copyright © Alescia Ann Roberto, 2011. All rights reserved

Page 3: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

ii

TABLE OF CONTENTS

LIST OF TABLES iv

LIST OF FIGURES v

ACKNOWLEDMENTS vi

ABBREVIATIONS viii

ABSTRACT ix

CHAPTER 1. INTRODUCTION

1.1 General Introduction 1

1.2 Aims and Outline of this Thesis 3

1.3 References 5

CHAPTER 2. LITERATURE REVIEW

2.1 Soil Microbial Function and Diversity 7

2.2 Biogeography of Microorganisms 8

2.3 The Soil Habitat: Influences on Microorganisms 9

2.3.1 Soil Chemistry: pH, Carbon and Nitrogen Dynamic 9

2.3.2 Soil Moisture 10

2.3.3 Landform and Soil Type 11

2.3.4 Land Management 11

2.3.5 Seasonality 12

2.4 Nitrogen Cycling in Soil 13

2.4.1 Denitrification 14

2.4.1.1 Cycling of Nitrogen (N) in the Denitrification Cycle 15

2.4.1.2 Human Alterations to the Denitrification Cycle and the

Consequences 16

2.5 References 17

CHAPTER 3. OPTIMIZATION OF QPCR ASSAY FOR MICROBIAL DNA

EXTRACTED FROM SOIL

3.1 Introduction 25

3.2 Methods, Results and Discussion

3.2.1 DNA Isolation 26

3.2.2 End Point PCR Optimization 26

3.2.3 Preparation of Stock I 27

3.2.4 Test Plate 27

3.2.5 Experiment 1: Comparison of Enzyme Performance Between

Brilliant II, PerfeCTa, Kappa, and Quantifast 28

3.2.6 Experiment 2: Effects of Annealing Temperature on Gene Target 29

3.2.7 Experiment 3: Field Site and Sampling Date Affect on Inhibition

Threshold. 30

Page 4: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

iii

3.2.8 Experiment 4: Effect of Annealing Temperature on Efficiency

and Sensitivity of Enzyme 31

3.2.9 Experiment 5: Effect of Slow Cycling with Quanta PerfeCTa

SYBR Green SuperMix 31

3.2.10 Experiment 6: Effect of Extended Denaturation and Additives

on qPCR Assay 32

3.2.11 Experiment 7: Extending Denaturation Time with Quanta 33

PerfeCTa FastMix and BSA

3.2.12 Experiment 8: Quanta PerfeCTa SuperMix versus FastMix

with BSA 33

3.2.13 Experiment 9: qnorB Small Amplicon Quanta FastMix

with BSA Summary 34

3.2.14 Experiment 11: nirS Test Plate with Quanta SuperMix 34

3.2.15 Experiment 12: Comparison of Touchdown Cycling

Conditions with nirS Gene Target between Quanta

SuperMix and Brilliant II 35

3.2.16 Experiment 13: Comparison of Non Touchdown Cycling

Conditions with nirS Gene Target between Quanta

SuperMix and Brilliant II 36

3.2.17 Experiment 14: Comparison between Touchdown Cycling

Conditions on Gene target nirS to no Touchdown Cycling 36

3.3 Conclusions 37

3.4 References 38

CHAPTER 4. SPATIO-TEMPORAL DISTRIBUTION OF QNORB-BEARING AND

NOSZ-BEARING DENITIRFIERS IN AND AGRO-ECOSYSTEM AND A PRAIRIE

ECOSYSTEM

4.1 Abstract 57

4.2 Introduction 57

4.3 Materials and Methods 61

4.3.1 Experimental site 61

4.3.2 Soil sampling 61

4.3.3 Physiochemical analysis 61

4.3.4 DNA Extraction 62

4.3.5 qPCR of qnorB and nosZ abundance 62

4.3.6 Statistical analysis 63

4.4 Results 63

4.4.1 Soil Series 64

4.4.2 Sampling Date 64

4.5 Discussion 65

4.6 Conclusion 69

4.7 Acknowledgments 70

4.8 References 70

APPENDIX DATA FOR CHAPTER 3 93

APPENDIX DATA FOR CHAPTER 4 100

Page 5: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

iv

LIST OF TABLES

CHAPTER 3

Table 3.1 Primers for molecular assays 41

Table 3.2 Eleven point dilution series for Stock I 41

Table 3.3 Master mixes and cycling conditions for Experiment 1: Comparisons between

enzymes 43

Table 3.4 Experiment 1: Enzyme repeatability and performance 45

Table 3.5 Master mix and cycling conditions for Experiment 2: Effect of annealing

temperature on gene target 45

Table 3.6 Gradient conditions for effect of annealing temperature on enzyme

performance 48

Table 3.7 Experiment 4: Annealing temperature affect on efficiency of Quanta enzyme 49

Table 3.8 Experiment 5: Performance of Quanta PerfeCTa SuperMix enzyme: within

and between plate repeatability and efficiency 49

Table 3.9 Experiment 6: Effect of additives, master mixes and cycling conditions 50

Table 3.10 A The effect of BSA with and without 5% glycerol on SuperMix enzyme

performance 51

Table 3.10 B The effect of T4 gene with and without 5% Glycerol on SuperMix enzyme

performance 52

Table 3.11 Experiment 7: Performance of Quanta PerfeCTa FastMix with long denaturation

time and BSA 53

Table 3.12 Experiment 12: Performance of Quanta SuperMix and Brilliant II with long

denaturation and touchdown thermal cycling conditions 55

Table 3.13 Experiment 13: Performance of Quanta SuperMix and Brilliant II with longer

denaturation and no touchdown thermal cycling conditions 56

CHAPTER 4

Table 4.1 All subsets regression results testing for the ability of edaphic

and biological factors to predict nosZ abundance 88

Table 4.2 Abundances of 16S rRNA, qnorB, and nosZ genes at each sampling date.

Significant differences indicated by different letters (P < 0.05) 88

Table 4.3 All subsets regression in identifying edaphic and biological factors predicting

nosZ and qnorB abundance by sampling date 89

Page 6: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

v

LIST OF FIGURES

CHAPTER 3

Figure 3.1 qPCR inhibitory behavior of Stock I solution at different dilutions. 42

The bracketed Ct results generated at LOG-linear-amplification-

capable sample dilutions. The circled Ct results are the inhibitory

dilution rage at higher template dilutions.

Figure 3.2 qPCR Ct results exhibiting a straight line at 42

LOG-linear-amplification-capable dilutions

Figure 3.3 qPCR inhibitory behavior of Stock I solution at different dilutions 44

with enzyme kits: Brilliant II (diamond), Kappa (square), Quanta

(triangle), and Qiagen (x). The bracketed Ct results generated at

LOG-linear-amplification-capable sample dilutions. The hooked

Ct results are the inhibitory dilution rage at higher template dilutions.

Figure 3.4 qPCR Ct results exhibiting a straight line at LOG-linear-amplification 44

-capable dilutions for enzyme kits Brilliant II, Kapa, Quanta, and

Qiangen

Figure 3.5 qPCR Ct results exhibiting a straight line at 46

LOG-linear-amplification-capable dilutions for enzyme kits

Brilliant II (diamond and triangle) and Quanta (square and x) at varying

annealing temperatures for A) 16S rDNA (50˚C and 60˚C), B) nosZ

(50˚C and 60˚C), and C) qnorB (50˚C and 55˚C)

Figure 3.6 qPCR inhibitory behavior of different Stock I 47

(Kalsow site in August (diamonds) and October (squares)

and Uthe site (triangles)) solutions for nosZ gene target at

different dilutions with Quanta enzyme kit. The bracketed

Ct results generated at LOG-linear-amplification-capable

sample dilutions. The hooked Ct results are the inhibitory dilution

rage at higher template dilutions.

Figure 3.7 qPCR Ct results exhibiting a straight line at 47

LOG-linear-amplification-capable dilutions for May, August,

and October Kalsow site stock I and Uthe stock I

Figure 3.8 qPCR Ct results exhibiting a straight line at 48

LOG-linear-amplification-capable dilutions stock I at

varying annealing temperatures for nosZ gene target

Figure 3.9 qPCR Ct results exhibiting a straight line at 53

LOG-linear-amplification-capable dilutions stock I with

extended denaturation time with Quanta PerfeCTa FastMiz with A) and

without BSA B) (400ng/ul).

Page 7: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

vi

Figure 3.10 qPCR Ct results exhibiting a straight line at 54

LOG-linear-amplification-capable dilutions stock I

from reactions comparing Quanta PerfeCTa FastMix (diamonds)

and SuperMix (squares) with BSA (400ng)

Figure 3.11 qPCR inhibitory behavior of Stock I solutions for nirS 54

gene target at different dilutions with Quanta enzyme kit.

The bracketed Ct results generated at LOG-linear-amplification-capable

sample dilutions. The hooked Ct results (circled) are the inhibitory

dilution rage at higher template dilutions.

Figure 3.12 Melt curves of nirS gene target with Brilliant II enzyme run 56

with non touchdown (A) and touchdown (b) cycling conditions

CHAPTER 4

Figure 4.1: nosZ abundance averaged for each soil type. Bars with the 90

same letter were not significantly different (P ≤ 0.05).

Error bars represent ± standard error.

Figure 4.2: Sampling date effects on (A & B) 16S rRNA and nosZ by 91

land management and C) on nosZ by landform. Bars accompanied

by the same letter between months were not significantly different

(ANOVA; α ≤ 0.05). Error bars represent ± standard error.

Figure 4.3: Sampling date effects on A) Cu , B) EOC, C) TN, D) WFPS, 92

E) C:N Means accompanied by the same letter were not significantly

different (ANOVA; α ≤ 0.05). Error bars represent ± standard error.

Page 8: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

vii

ACKNOWLEDMENTS

I would like to thank my committee members Dr. Kirsten Hofmockel, Dr. Larry Halverson, and

Dr. Thomas Loynachan. The advice, criticisms, and encouragements from my advisory

committee members and lab mates Elizabeth Bach, Sarah Hargreaves, and Ryan Williams were

much appreciated. I would also like to give special thanks to my sampling crew: Steven Adams,

Sarah, Dr. Halverson, Matthew Raymond, Seth Wenner, and Stith Wiggs. This work was made

possible by funding from the department of Ecology, Evolution, and Organismal Biology.

My final gratitude’s are to my family: Frank, Diane, Danielle and Christina, my boyfriend,

Steve, and lab mate Sarah. This would not be possible without their support and

encouragements…… I am forever grateful to have all of you in my life.

Page 9: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

viii

ABBREVIATIONS

qnorB gene encoding the quinol membrane-bound nitric oxide reductase

nosZ gene encoding the nitrous oxide reductase

nir-type gene encoding the copper-containing or cytochrome cd1 nitrite reductase

PCR polymerase chain reaction

qPCR quantitative polymerase chain reaction

N2 dinitrogen gas

N2O nitrous oxide

NH4+ ammonium

NO3-

nitrate

NO nitric oxide

WFPS water-filled pore space

C: N ratio carbon to nitrogen ratio

Fe iron

Cu copper

EOC extractable organic carbon

TC total carbon

TN total nitrogen

pH soil pH

Page 10: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

ix

ABSTRACT

Denitrification exhibits high spatial and temporal variability (Nunan et al., 2002), and, therefore,

is difficult to predict. Since denitrifiers mediate denitrification, understanding the factors that

influence the distribution of these microbes may help explain the variation observed in process

rates. In this work I examined qnorB-bearing and nosZ-bearing denitrifiers and total bacteria, in

terms of abundance in different soils. Denitrifying bacteria were described using genes that

encode for the enzymes responsible for the production and reduction of N2O. Appropriate targets

and methods for this were evaluated. An assessment was then carried out of how the distributions

of these biological factors are affected by spatio-temporal differences of soil properties.

The seasonal variation in qnorB and nosZ abundance indicates a change in the genetic potential

of the denitrifier community’s ability to respond to fluctuations in soil resources. However, the

abundance of nosZ only correlated to the inherent properties at landform scale, with unique

parameters related to nosZ-bearing denitrifier being micronutrients and qnorB abundance. Of the

soil properties measured, none consistently explained the pattern observed in gene abundance

through time or space, except nitrogen and micronutrients. In addition, qnorB abundance did not

correlate with the environmental gradients measured in this study, suggesting that other forms of

physiological cues influence the distribution of this gene. These findings suggest that the

variation detected in qnorB and nosZ abundance is dependent on physiological responces to

either oxygen concentration or essential nutrients such as trace metals.

Page 11: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

1

1

CHAPTER 1. INTRODUCTION

1.1 General Introduction

Microorganisms are the foundation of life. They are the drivers of biogeochemical processes,

such as denitrification, and are solely responsible for the cycling of many nutrients in

ecosystems. Therefore, when the importance of environmental health became a concern to

mankind, understanding the scope and significance of microbial involvement in ecosystem

processes has become increasingly important to the scientific community. Indeed, as the

pressures on land, chiefly the rising demand for shelter and food production, have caused and

expedited the reduction of our earth’s land, water, and air, researchers have recently begun to

focus on microorganisms due to their importance in the role of maintaining soil fertility, soil

structure, water purity, and nutrient availability and cycling. Unfortunately, only about 1% of

these organisms can be cultured and even fewer are known to have specific soil functions

(Amann et al., 1995).

Terrestrial composition of microorganisms is highly diverse and not easily observed. Because

soil bacteria carry out fundamental ecosystem processes, understanding the link between

diversity and function in soil is of importance, specifically in the development of sustainable

ecosystem management (Enwall et al., 2010). With the advent of molecular techniques in

microbial ecology, we have been able to employ trait-based studies to survey the biogeography

of microorganisms. Evidence from molecular research has made us increasingly aware of the

significance of specific microorganisms in nature. One such group of organisms, denitrifiers, is

vital due to their chief function in cycling of nitrogen from nitrite and nitrogenous oxides to inert

nitrogen gas.

Denitrification is a facultative process in which oxidized forms of nitrogen are used to support

anaerobic respiration in an oxygen-limited environment. The genes responsible for the

conversion of nitrogen oxides to inert nitrogen gas belong to diverse groups of unrelated and

phylogenetically different bacteria (Zumft, 1997; Mergel et al., 2001; Van Spanning, 2005) For

this reason, 16S rDNA rDNA, a molecular marker commonly used for analyzing the community

structure of bacteria, is not suitable for denitrifier community composition analysis; instead the

genes associated with the denitrification process are used.

Page 12: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

2

Not all denitrifying bacteria produce the complete suite of genes encoding enzymes to complete

the pathway (Zumft, 1997). As a result, the byproducts of an incomplete pathway, specifically

nitric (NO) and nitrous (N2O) oxides, are produced during the dissimilatory reduction process are

a major concern. These nitrogenous oxide species not only reduce growth and yields of cropped

plants, but also are partially responsible for the depletion in water quality and of the ozone layer,

and for the initiation of acidic rains and global warming. Therefore, the genes associated with

denitrification, specifically in the reduction of nitric (NO) and nitrous (N2O) oxides, nitric oxide

reductase (norB) and nitrous oxide reductase (nosZ), will be used in this thesis to study the

ecology of the specific guilds of bacteria possessing these genes.

The “prairie pothole region” encompasses the area from Alberta, Canada, to Iowa, USA, is the

largest wetland habitat in North America, and consists of temporary, seasonal, and permanent

wetlands within prairie (Euliss et al., 1999). These wetlands or “potholes” are depressions

connected by shallow drainage ways formed approximately 13,000 years ago by the recession of

glaciers from the Des Moines Lobe Wisconsin Glacial event (Euliss et al., 1999). The positive

N2O fluxes associated with this area, are often characterized by topography (e.g., water

accumulation in depressions), initiated by events of rain or snow-melt (Dunmola et al., 2009;

Gleason et al., 2009; Ma et al., 2008; Yates et al., 2006b). It is under these conditions that

denitrification dominates (Corre et al., 1996).

Microbial community composition is one of the most significant influences on both the turnover

of solutes in soil and water and global trace gas emissions in denitrification (Holtan-Hartwig et

al., 2000), however theses activities are highly dependent upon soil characteristics. The available

oxygen (O2) (Cavigelli and Robertson, 2000), C, and N (Avrahami et at., 2002; Svensson et al.,

1991; Wang et al., 2005), along with land management and landform (soil type), have been

shown to influence denitrification process rates (Enwall et al., 2010; Girvan et al., 2003;

Steenwerth et al., 2002). Though land management and landform are distal controls (Wallenstein

et al., 2006), land management mainly influences nutrient availability (crop cover and nitrogen

additions) and soil disturbance (tillage) (Bruns et al., 1999; Stress et al., 2004), whereas landform

affects water and nutrient distribution (Yates et al., 2006a; Yates et al., 2006b). Together, these

factors influence microbial community composition and abundance, and thus reflect long-term

climate, disturbance, and resource availability of soils (Cavigelli and Robertson, 2000).

Page 13: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

3

1.2 Aims and outline of thesis

The main aim of this thesis was to study the abundance of genes encoding a portion of the

enzymes for the denitrification pathway in one of Iowa’s prairie pothole systems. In this work,

the quantitative polymerase chain reaction (qPCR) molecular method was used to study the size

of gene pools of the bacterial denitrifier community in natural prairie and agriculture soil. The

primer sets amplifying qnorB and nosZ genes have been published and used previously in

environmental surveys of denitrifiers (Braker et al., 1998, 2010; Henry et al., 2006). However,

the primers targeting the gene of interest must be both specific and sufficiently sensitive.

Therefore, we started by evaluating how soil characteristics, sampling time, enzyme chemistries,

and cycling conditions interacted to influence both qPCR efficiencies and sensitivity. Once

reactions were optimized, we explored the biogeography of denitrifying communities. The main

objectives of this study were to determine the extent to which 1) qnorB and nosZ gene

abundances vary spatially, 2) qnorB and nosZ gene abundances vary temporally, and 3) to

determine the environmental factors associated with landform, land management, and sampling

date that explain variation of denitrifier abundance. To answer these questions, qnorB-bearing

and nosZ-bearing denitrifier abundances were determined for two common landscapes

characteristic of the North American “prairie pothole region”: cultivated potholes and

uncultivated (prairie) potholes. These questions were answered by a study designed to test the

following hypotheses:

1) abundance of qnorB and nosZ will be greatest in low-lying sites, where nutrient

availability (C and N) and water concentrations are greatest

2) abundance of qnorB and nosZ will be greater in the prairie due to greater nutrient

availability and wetter conditions

3) abundance of qnorB and nosZ will fluctuate with sampling date and will correlate with

changes in abiotic soil properties and

4) abundance of qnorB and nosZ will be influenced by the concentrations of enzymatic

co-factors associated with the enzymatic nitric oxide reductase and nitrous oxide

reductase.

To accomplish this, an observational study was done at the Kalsow Prairie field site. The site

contains a 160-acre tall grass prairie state preserve located in southern Pocahontas County,

Iowa, U.S.A., and a cultivated corn-corn-soybean rotated field located directly south and

Page 14: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

4

adjacent to the prairie (within 100m).

Both sites are located on Iowa’s youngest geological landform, experiencing its last glacial

event about 13,000 years ago (Prior, 1991), and contain potholes typical of the Des Moines Lobe

Landform region. Kalsow Prairie is one of two original prairie preserves purchased by the state

in 1949 (Dornbush), representing one of the five largest virgin prairies remaining within the

glaciated region of Iowa (Rosburg, 2001). The cropped field site has been in yearly corn-corn-

soybean rotation for the past 13 years, and is under a conventional management regime.

As in all of Iowa, Pocahontas County’s weather is characterized by marked seasonal variations.

These variations result from a moist southerly flow, which originates in the Gulf of Mexico and

produces the majority of precipitation during the summer, and a northwesterly flow, which

originates in Canada and produces the majority of the precipitation during the winter.

Throughout the summer, air masses periodically shift and air from the Pacific Ocean brings mild,

dry weather as well as hot, dry winds from the Southwest resulting in high temperatures and crop

desiccation (NCDC). IEM Climodat has collected precipitation and temperature data for

Pocahontas County for the past 57 years (1951-2008) and reports show the average precipitation

for the county to be 30.34 in over that time. More specifically, average spring precipitation

(March-May) is 8.84 in, summer precipitation (June-August) is 12.62 in, fall precipitation

(September-November) is 6.42 in, and winter precipitation (December-February) is 2.46 in. The

annual mean temperature ranges from -8.8˚C in January to 22.7˚C in July (Carlson and Todey,

2009).

USDA soil surveys (Soil Survey) have characterized four distinct surface horizon soils in the

pothole region of Southern Pocahontas County; Clarion (summit), Webster (side-slope), Canisteo

(foot-slope), and Okoboji (closed depression). The upland areas, dominated by Webster and

Clarion soils, are characterized as a fine-loamy, mixed, mesic Typic Endoaquolls and

Haplaquolls (Finley and Lucassen, 1985). The lowland areas, dominated by Canisteo and

Okoboji soils, are characterized as fine-loamy, mixed, superactive, calcareous, mesic Typic

Endoaquolls and fine, smectitic, mesic Cumulic Vertic Endoaquolls (NRCS).

Page 15: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

5

1.3 References

Avrahami, S., R. Conrad, et al., (2002). "Effect of soil ammonium concentration on N2O release

and on the community structure of ammonia oxidizers and denitrifiers." Applied and

Environmental Microbiology 68(11): 5685-5692

Bruns, M. A., J. R. Stephen, et al., (1999). "Comparative Diversity of Ammonia Oxidizer 16S

rRNA Gene Sequences in Native, Tilled, and Successional Soils." Appl. Environ.

Microbiol. 65(7): 2994-3000.

Carlson, R. E. Aad D. Todey. 2009. Iowa State University Climatological data from 1951 to

2000. Data averaged from meteorological stations located in Pocahontas and Rockwell-

City, Iowa, U.S.A. <http://mesonet.agron.iastate.edu/climodat/>.

Cavigelli, M. A. and G. P. Robertson. (2000). "The functional significance of denitrifier

community composition in a terrestrial ecosystem." Ecology 81(5): 1402-1414.

Dunmola, A., M. Tenuta, et al., "Pattern of greenhouse gas emission from a Prairie Pothole

agricultural landscape in Manitoba, Canada." Canadian journal of soil science 90(2): 243-

256.

Enwall, K., I. N. Throback, et al., (2010). "Soil Resources Influence Spatial Patterns of

Denitrifying Communities at Scales Compatible with Land Management." Applied and

Environmental Microbiology 76(7): 2243-2250.

Euliss, N. H., Jr., David M. Mushet, and Dale A. Wrubleski. (1999). "Wetlands of the Prairie

Pothole Region: Invertebrate Species Composition, Ecology, and Management."

Invertebrates in Freshwater Wetlands of North America: Ecology and Management,

Chapter 21 Retrieved 4/1/2009, 2009, from

http://www.npwrc.usgs.gov/resource/wetlands/pothole/index.htm (Version 02SEP99).

Finley, R. D. and J. A. Lucassen. (1985). "Soil Survey of Pocahontas County, Iowa." Soil Survey

of Pocahontas County, Iowa.: 128 pp.

Girvan, M. S., J. Bullimore, et al., (2003). "Soil type is the primary determinant of the

composition of the total and active bacterial communities in arable soils." Applied and

Environmental Microbiology 69(3): 1800-1809.

Gleason, R. A., B. A. Tangen, et al., (2009). "Greenhouse gas flux from cropland and restored

wetlands in the Prairie Pothole Region." Soil Biology and Biochemistry 41(12): 2501-

2507.

Holtan-Hartwig, L., P. Dorsch, et al., (2000). "Comparison of denitrifying communities in

organic soils: kinetics of NO3- and N2O reduction." Soil Biology & Biochemistry 32(6):

833-843.

Ma, W. K., A. Bedard-Haughn, et al., (2008). "Relationship between nitrifier and denitrifier

community composition and abundance in predicting nitrous oxide emissions from

ephemeral wetland soils." Soil Biology and Biochemistry 40(5): 1114-1123.

Page 16: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

6

Mergel, A., K. Kloos, et al., (2001). "Seasonal fluctuations in the population of denitrifying and

N-2-fixing bacteria in an acid soil of a Norway spruce forest." Plant and Soil 230(1): 145-

160.

Soil Survey Staff, Natural Resources Conservation Service, United States Department of

Agriculture. Web Soil Survey. Available online at http://websoilsurvey.nrcs.usda.gov/

accessed [2009].

Steenwerth, K. L., L. E. Jackson, et al., (2002). "Soil microbial community composition and land

use history in cultivated and grassland ecosystems of coastal California." Soil Biology

and Biochemistry 34(11): 1599-1611.

Stres, B., I. Mahne, et al., (2004). "Nitrous oxide reductase (nosZ) gene fragments differ between

native and cultivated Michigan soils." Applied and Environmental Microbiology 70(1):

301-309.

Svensson, B. H., L. Klemedtsson, et al., (1991). "Soil denitrification in three cropping systems

characterized by differences in nitrogen and carbon supply." Plant and Soil 138(2): 257-

271.

Van Spanning, R. J., Delgado, M. J., and Richardson, D. J. (2005). The nitrogen cycle:

Denitrification and its relationship to N2 fixation. Nitrogen Fixation in Agriculture,

Forestry, Ecology and the Environmental. a. W. E. N. D.Werner, eds. Springer

Netherlands, Dordrecht, The Netherlands. 4: 277-342.

Wallenstein, M. D., D. D. Myrold, et al., (2006). "Environmental controls on denitrifying

communities and denitrification rates: Insights from molecular methods." Ecological

Applications 16(6): 2143-2152.

Wang, L., Z. Cai, et al., (2005). "Effects of disturbance and glucose addition on nitrous oxide

and carbon dioxide emissions from a paddy soil." Soil and Tillage Research 82(2): 185-

194.

Yates, T. T., B. C. Si, et al., (2006). "Probability distribution and spatial dependence of nitrous

oxide emission: Temporal change in hummocky terrain." Soil Science Society of

America Journal 70(3): 753-762.

Yates, T. T., et al., (2006). Wavelet spectra of nitrous oxide emission from hummocky terrain

during spring snowmelt. Madison, WI, ETATS-UNIS, Soil Science Society of America.

Zumft, W. G. (1997). "Cell biology and molecular basis of denitrification." Microbiology and

Molecular Biology Reviews 61(4): 533.

Page 17: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

7

CHAPTER 2. LITTERATURE REVIEW

2.1 Soil Microbial Function and Diversity

Soil bacteria play a central role in ecosystem functioning via biogeochemical processes, and are

responsible for the transformation of carbon, nitrogen, phosphorus, and sulfur. More than 90% of

the energy in a soil system passes through the microbial population (Nannipieri and Badalucco,

2003), in part because bacteria are ubiquitous, largely diverse, both in genetics and in richness

(the number of different species in a given area), and have a high range of metabolic activities

(Paul, 2007). Billions of organisms and thousands of species are known to be found in one gram

of soil (Torsvik et al., 1990), yet less than one percent of microbial species are capable of being

cultivated and characterized (Amann et al., 1995). Furthermore, because most organisms have

yet to be identified and characterized, the basis of functional groups is highly dependent on the

knowledge of a few species (Nannipieri and Badalucco, 2003).

In addition, process level measurements are thought to be insensitive to community level

changes since functional groups and the processes they perform are thought to be functionally

redundant, i.e., more than one species performs a similar function in an ecosystem process, and

therefore biodiversity does not have a major impact on ecosystem functioning (Beare et al.,

1995; Yin et al., 2000). Given that not all species interact in the same way with other species,

functional redundancy is difficult to define and involves many different niche dimensions

(Rosenfeld, 2002). Moreover, microbes vary in the contributions made to the processes they

perform (e.g., rates of processes and pathways utilized to process resources), and hence a loss of

a “species” may not be completely compensated by a functionally similar species (Chapin et al.,

1997; Orians, 1997).

Studies have demonstrated that soil microbial communities can change in response to

disturbances (Atlas et al., 1991; McCaig et al., 1999), and these disturbances alter ecosystem

processes (Freckman et al., 1997), which in turn modify community composition (Chapin et al.,

1997). Depending on the organisms and the traits they possess, environmental changes could

either stabilize the functionality of an ecosystem or trigger dramatic functional changes. This is

because in heterogeneous environments selection via resource partitioning, such as the use of

metabolic by-products, allows for the evolutionary emergence of organisms capable of co-

existing, resulting in niche complementarity (Rainey et al., 2000; Travisano and Rainey 2000).

Page 18: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

8

This diversity is important in maintaining complex biotic interactions, which determine

biogeochemical cycling and spacio-temporal heterogeneity (Beare et al., 1995), and ultimately

contributes to the productivity and resilience of soil ecosystems (Torsvik and Øvreås, 2002;

Yachi and Loreau, 1999). Hence, one cannot predict how ecosystem functions will change until

one understands not only the responses of microbial communities in relation to the individuals

that comprise that community, but also how these individuals and communities fluctuate within

their habitat (Kennedy, 1999; O’Donnell et al., 2001).

2.2 Biogeography of Microorganisms

To determine spatial distribution of microorganisms, information is needed about the abundance

of organisms and how their environment “selects” or influences the spatial variation in microbial

diversity (Martiny and Bohannan, 2006). However, both environmental and evolutionary factors

affecting microbial distribution are often difficult to identify and are not likely to be comparable

in every ecosystem. These setbacks are often overcome by biogeographical studies, which can

provide insight into the environmental and historical pressures shaping the community of interest

(Cho and Tiedje 2000; Bohannan and Hughes 2003; Fierer and Jackson 2006; Martiny,

Bohannan et al. 2006; Prosser, Bohannan et al. 2007). Although biogeography can span various

scales, site-specific or local scales have advantages over large-scale studies. One of these

advantages is that environmental factors influencing microbial community composition and

diversity are likely to be more similar than in large-scale studies (Cho and Tiedje 2000; Green

and Bohannan 2006; Martiny, Bohannan et al. 2006). Additionally, systems such as the prairie

pothole ecosystem, that provide natural gradients, allow for comparative biogeographical studies

designed to facilitate our understanding of how environmental factors and soil structure influence

community abundance and composition over space and time (Martiny, Bohannan et al. 2006).

Thus, they provide insight into how microbial community abundance might play a role in

determining ecosystem process rates.

Multiple forces are known to structure microbial communities and their abundances; microbial

community structure, composition, and abundance in a given time or space are most likely the

result of interaction effects of the habitat and the community of interest (Martiny and Bohannan,

2006). More often, reports have focused on the total bacterial community (Franklin and Mills,

2003; Green et al., 2008), with very few studies looking at functional traits (Enwall et al., 2010;

Page 19: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

9

Philippot et al., 2009). Because very little is known about the spatial distribution of the microbial

functional traits that mediate ecosystem processes, characterization of spatial patterns of these

traits is needed to identify how edaphic characteristics mediate resource-based biogeographical

distribution. Understanding what mediates patterns of distributions in microbial communities

would not only provide an insight into the underlying mechanisms causing temporal and spatial

variation but also facilitate our understanding of microbial community ecology and ecosystem

processes.

2.3 The Soil Habitat: Influence on Microbial Communities

Soil is a complex, heterogeneous and discontinuous environment that supports a complex

community of microorganisms that play a critical role on Earth (Paul, 2007; Nannipieri and

Badalucco, 2003). Variation of soil properties can span various scales, which differ in chemical,

physical, and biological characteristics over both space and time. These characteristics, pH,

water-filled pore space (O2 availability), carbon, and nitrogen availability, all of which change

with season, soil type, and land management, contribute to the heterogeneity and complexity of

microbial habitats and ultimately affect the ecology, activity, and population dynamics of soil

microorganisms (Baath et al., 1998; Fierer et al., 2003; Holland and Coleman, 1987; Marschner

et al., 2003; Nannipieri and Badalucco, 2003; Torsvik et al., 1996). Therefore, microorganisms,

specifically bacteria, are directly affected by both the physical location and the physical makeup

of their habitat (Paul, 2007; Stotzky, 1997). Understanding how these physical aspects influence

microbial communities will provide insight in predicting ecosystem processes and behavior

(Swanson et al., 1988).

2.3.1 Soil Chemistry: pH, Carbon and Nitrogen Dynamics

Soil chemistry or the biochemistry (chemical reaction, activities and products of soil

microorganisms) of soils is determined mainly by soil pH and redox reactions. Proton and

electron exchanges through soil solutions define the environment controlling microbial activity

(Paul, 2007). These exchanges are highly dependent on electron availability, which in turn is

influenced by the effect of soil pH on the solubility and ionization of inorganic and organic soil

components. In terms of energy, however, microbial activity has a significant effect on

regulating the flow of electrons and protons though the soil. The generation of electrons occurs

Page 20: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

10

through metabolic oxidative reactions of substrates, with the exchange of electrons from donors

(soil organic matter (SOM) to acceptors (oxygen, nitrogen oxide, etc.).

Inputs of SOM are usually from plant residues, decaying organisms, or manure, and vary in its

physical characteristics, chemical composition, biological availability, and amounts assessable to

heterotrophic microorganisms (Baggs et al., 2003; Chaves et al., 2005; Wang et al., 2005).

Organic matter is an important energy source for soil microorganisms and a primary source of

mineralizable nitrogen, sulfur, and phosphorus; it influences the availability of metal ions in soils

by forming soluble complexes (Stevenson, 1994). An increase in SOM enhances biological

oxygen demand for aerobic respiration, and respiration and substrate diffusion rates control the

availability of electron acceptors. If the diffusion of oxygen (the primary electron acceptor in

respiration) cannot satisfy its consumption (Klemedtsson et al., 1991) due to water-logging or

restricted pore size due to soil texture or compaction, soils become anoxic and microbial activity

is then controlled by the movement of electrons to other available alternative acceptors, thus,

allowing for the reduction of elements such as nitrogen, iron, sulfur, and carbon dioxide in soils.

The availability of both carbon and alternative acceptors drives the rate at which these elements

are being utilized. For example, NO3- is a preferred acceptor of electrons over other nitrogenous

species, such as N2O; hence N2O will accumulate when NO3- supply exceeds its demand (Swerts

et al., 1996).

2.3.2 Soil Moisture

To some extent, rates of mineralization, nitrification, and denitrification are controlled by the

transport and diffusion of solutes and gases in water and air. Soil water status influences process

rates largely due to the rate of gas or substrate diffusion (Farquharson and Baldock, 2008). Influx

of moisture, specifically after a rain event, allows for water bridges to form between various soil

particles and aggregates, allowing for movement of substrates and organisms. In water-logged

soils, diffusion rates of oxygen decrease and oxygen is often used by microorganisms faster than

it is replenished (Klemedtsson et al., 1991). Rates of nutrient use occur most rapidly in the

presence of oxygen, and more slowly for other electron acceptors. Hence, soil water content

directly influences both oxygen availability and the use of alternative electron acceptors. Also of

Page 21: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

11

note, the distribution and volume of water in the soil are mainly determined by the physical

structure of the soil and topography of the land (Yu et al., 2001).

2.3.3 Landform and Soil Type

Landforms regulate movement of materials (nutrients and sediments), organisms and energy by

defining gravitational gradients (Swanson et al., 1988). The role of landforms is important in

studies of nutrient cycling because spatial patterns of bigeochemical process fluxes (e.g.

denitrification) are influenced by the movement of water and material (Swanson et al., 1988).

However, parent material (i.e., soil type) also plays an important role influencing water and

nutrient availability (Yu et al., 2001). This is due in part to the ability of the soil to absorb

molecules, such as organic and inorganic compounds, which reduces substrate accessibility to

soil microorganisms (Ladd et al., 1996). Soil structure and aggregation, which are dependent on

soil type, influence the rate of diffusion of solutes and gases through the soil (Farquharson and

Baldock, 2008) by mediating the water films that surround the pores (Nunan et al., 2003), thus

altering microbial activity and community structure at the field scale (Linn and Doran, 1984).

In the subtle topography of catenas, for example, higher concentrations of organic carbon,

nitrogen, and phosphorus were observed to increase going down the catena (Schimel et al.,.,,

1985). Furthermore, lower sloped position soils are more fertile and contain greater fine soil

particles, such as silt and clay, as a result of subsurface flow and, therefore, have greater

moisture holding capacities (Swanson et al., 1988). This deposition of nutrients and water has

been associated with higher rates of ecosystem processes and microbial activity in lower lying

soils (Hanson et al., 1994).

Consequently, most landforms have been influenced by human land use, and those influenced by

anthropogenic disturbances may influence microbial abundances differently than natural

landforms.

2.3.4 Land Management

Land management differences, such as grasslands versus cultivated soil, distinctly influence soil

characteristics and soil microbiology (Steenwerth et al., 2002). Grassland soils support and

maintain increased levels of soil carbon, microbial biomass, and greater spatial heterogeneity of

a soil profile (Calderón et al., 2000; Kandeler and Murer, 1993) as compared to cultivated soil

Page 22: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

12

that has been subjected to repeated tillage (Burke et al.,1989; Conant et al., 2001; Schimel et al.,

1985). Additionally, anthropogenic nutrient inputs, such as fertilizers and pesticides, along with

residue incorporation, cropping sequence, and tillage have been shown to alter both nutrient

resources and microbial biomass and composition (Anderson and Gray, 1990; Calderón et al.,

2000; Enwall et al., 2005, 2007, 2010; Miller et al., 2008; Patra et al., 2006). Additionally,

studies using nucleic acid methodology have demonstrated differences between microbial

communities from different management regimes (McCaig et al., 1999)

2.3.5 Seasonality

Seasonal changes in precipitation and temperature control the input and loss of water from soil

over time. Seasonal changes in weather control both soil moisture status and C and N availability

(Groffman et al., 2000). An example would be the increase in soil organic matter across a prairie

region as mean annual precipitation increases. Furthermore, respiration rates, assumed to be a

measure of the general soil microbial activity (Anderson, 1982), are highly dependent on

temperature and the availability of nutrients. Rises in temperature are known to cause an increase

in substrate availability, mineralization rates of soil organic matter, and decomposition rates via

increasing diffusion of soluble substrates (Davidson and Janssens, 2006), and to also increase the

rate of physiochemical reactions by decreasing the energy barrier for reactions (Farquharson and

Baldock, 2008). Very few soils, however, maintain a uniform temperature in their surface layers,

resulting in fluctuation either diurnally or seasonally. This fluctuation influences microbial

activity by affecting the community structure and composition (Braker et al., 2010; Farquharson

and Baldock, 2008; Szukics et al., 2010). The response of microbial processes to temperature has

an optima, which may occur due to energy tradeoffs (Stark and Firestone, 1996), and has been

found to occur at temperatures experienced in the soil at a particular site (Breuer and Butterbach-

Bahl, 2005; Farquharson and Baldock, 2008).

As temperature rises, enzymatic reactions increase, allowing for greater energy production; this

excess energy may be required for maintenance of enzymes and other cellular components that

either denature or become damaged at higher temperatures (Farquharson and Baldock, 2008;

Stark and Firestone, 1996). Thus, microbial mediated processes adapt to the climate at a

particular site (Powlson et al., 1988; Stark and Firestone, 1996), allowing for net energy

production. Evidence for temperature adaptation has been demonstrated for both denitrification

Page 23: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

13

and nitrification (Powlson et al., 1998; Stark and Firestone, 1996). When comparing the response

of denitrification to temperature in a temperate English soil and subtropical Australian soil, for

example, Powlson et al., (1988) found that denitrifiers from the English soil were better adapted

to denitrify at lower temperatures. Similarly, Stark and Firestone (1996) found that soil taken

from the open sites between oak trees had a higher temperature optimum for nitrification than

soils taken from under the canopy of trees within the same site where soil temperatures were

lower. Adaptation of microbial communities to temperature changes may either be physiological

or shifts in community structure to fluctuations of soil resources.

2.4 Nitrogen (N) Cycling in Soil

Nitrogen (N) is an essential limiting element required for the synthesis of many of the

biomolecules that all organisms need. In terrestrial and aquatic ecosystems nitrogen is found in

many oxidation states, the most important of which are nitrate, nitrite, nitric oxide, nitrous oxide,

dinitrogen, and ammonium (Van Spanning et al., 2005). Production and consumption rates of

bacterial metabolic processes are the main determinants of the concentrations of these

compounds as a free state and the main driving force of the global nitrogen cycle (Van Spanning

et al., 2005). Most nitrogen, approximately 78% of the earth’s atmosphere, is found in the

biologically inert state of nitrogen gas, a compound that consists of two N atoms joined via a

triple bond. A great deal of energy (945 kJ/mol) must be derived through atmospheric, industrial,

or a biological process to break this bond (Silberberg).

Atmospheric nitrogen fixation occurs via lightning or fires. The high temperatures created by

these natural phenomena result in a reaction between dinitrogen gas (N2) and oxygen (O2)

creating nitric oxide (NO). Once formed, NO interacts with rain, resulting in the formation of

nitric acid which is seeded into terrestrial and aquatic ecosystems as aqueous nitrate (NO3-).

Atmospheric fixation contributes 5-8% of the total nitrogen fixed per year (Gruber 2008). In

addition, industrial nitrogen fixation occurs through several processes. However, the Haber-

Bosch process, an industrial process that combines N2 and hydrogen (H) under high pressure and

heat to produce ammonia (NH3), is the leading producer of biologically reactive nitrogen (160Tg

N y-1

), surpassing the level of nitrogen fixation both on land (110Tg N y-1

) or in the ocean

(140Tg N y-1

) (Gruber 2008). Lastly, biological N fixation (BNF) is the reduction of

atmospheric dinitrogen to ammonia (NH3). The reaction is carried out by the enzyme

Page 24: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

14

nitrogenase, which is only found in nitrogen fixing bacteria (diazotrophs) and/or archaea. BNF

contributes about 25 kg N ha-1

y-1

(Bacon, 1995).

Only when N2 is “fixed,” either through atmospheric, industrial or biological nitrogen fixation, is

it usable for other organisms. Once fixed, the available nitrogen becomes incorporated into

biomass either by free-living nitrogen fixing bacteria or the plants that are in a symbiotic

relationship with the nitrogen-fixing bacteria. It is only through the metabolic processes of

assimilation and decomposition that organic nitrogen, that is, nitrogen that was incorporated into

biomass, is converted to inorganic nitrogen. This conversion process, known as mineralization,

occurs when microorganisms release nitrogen as a byproduct of decomposition of organic matter

(OM). Mineralization occurs if the ratio of carbon to nitrogen (C: N) in OM is low (20:1),

meaning nitrogen is not limiting (Paul and Clark, 1989). If soil OM has a high C: N ratio,

nitrogen is instead used in growth production, a process known as immobilization.

Mineralized nitrogen (NH3) is readily oxidized to ammonium (NH4+) in terrestrial ecosystems. In

an aerobic environment, sequential steps of biological oxidation, performed by nitrifying bacteria

and archaea, oxidizes NH4+ to nitrite (NO2

-) and then to nitrate (NO3

-). Once nitrogen is

converted into NO3-, it is a usable form for a variety of different organisms and can be easily

mobilized by water (Brady and Weil, 2002). To maintain a balance between inert and usable

sources of nitrogen, a cycle must be formed to return this essential element back to its original

state. It is only through denitrification that nitrogen introduced into the biosphere is removed via

the production of atmospheric nitrogen (Zumft, 1997).

2.4.1 Denitrification

One of the main branches of the global nitrogen cycle is denitrification. This process is mainly

sustained by denitrifers, microorganisms that play a critical role in regulating N cycling through

the transportation of nitrogen from terrestrial ecosystems into the atmosphere. Though nitrogen

is crucial to the biota found on Earth, its role in the global denitrification cycle is also significant

to the functioning of the Earth itself, specifically, the reduction of nitrogenous oxides (NO and

N2O) to inert N2 gas (Zumft, 1997). In consequence, the byproducts of an incomplete

denitrification cycle have mainly adverse effects on the atmosphere, soil and water, resulting in

agronomic, environmental and human impacts.

Page 25: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

15

2.4.1.1 Cycling of Nitrogen (N) in the Denitrification Cycle

Denitrification, a process responsible for the transformation of terrestrial nitrogen to atmospheric

nitrogen, is an anaerobic respiratory process carried out by many bacterial and fungal species.

Nitrogen oxides are substituted for oxygen (O2) as a terminal electron acceptor resulting in the

process of dissimilatory nitrate reduction through a series of gaseous nitrogen intermediates.

Hence, O2 availability limits denitrification (Bateman and Baggs, 2005; Cavigelli and Robertson,

200). In bacteria this process proceeds through a series of four reduction reactions under which

nitrate (NO3-) is reduced to the intermediates nitrite (NO2

-), nitric oxide (NO), and nitrous oxide

(N2O) in successive steps to produce nitrogen gas (Ferguson, 1994; Stouthamer, 1991, 1992;

Cavigelli, 2000). The nature of the enzymes catalyzing these reactions belongs to diverse groups

of unrelated and phylogenetically different bacteria (Zumft, 1997; Mergel, 2001; Van Spanning,

2005). However, not all denitrifying bacteria produce the complete suite of enzymes to complete

the pathway (Zumft, 1997), and denitrifcation can thus be considered a community process.

The reduction of nitrate occurs by either a membrane bound nitrate reductase (Nar) and or a

peripalsmic one (Nap) Both receive electrons from quinol and catalyse the reduction of nitrate

(NO3-) to nitrite (NO2

-). Additionally both are found in both denitrifiers and non-denitrifiers (Van

Spanning et al., 2005).

Nitrite reductase is the key enzyme responsible for catalyzing nitrite (NO2-) to nitric oxide (NO),

the first committed step that produces gaseous intermediates in the denitrification cycle (Zumft,

1997). There are two types of nitrite reductases characterized in the form of copper (Cu)-Nir or

cadmium (Cd)-Nir. Both Cu-Nir (nirK), a copper-containing nitrite reductase, and Cd-Nir (nirS),

a cytochrome cd1 nitrite reductase (Zumft, 1997), are functionally equivalent periplasmic

proteins (Van Spanning et al., 2005; Jones et al., 2008). No organism has yet been found with

genes encoding for both enzymes, but some have been observed to have multiple copies nirS

(Etchebere and Tiedje, 2005).

NO reduction to N2O is catalyzed by nitric oxide reductase (Nor), a transmembrane, metallo-

enzyme with a heme active site. The enzyme is encoded by norB. Two types of Nor genes, cnorB

and qnorB, have been identified. Enzymes encoded by cnor receive electrons from either

Page 26: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

16

cytochrome c or pseudoazurin, whereas qnorB-encoded enzymes receive electrons from a quinol

pool (Zumft, 1997).

The final step in converting terrestrial nitrogen to atmospheric nitrogen is the reduction of N2O

to N2, catalyzed by nitrous oxide reductase. Nos is a periplasmic homodimeric enzyme with a

multi-copper active site. The enzyme is encoded by the gene nosZ, and due to its location in the

bacterial membrane it is sensitive to changes in the environment (Firestone et al., 1980; Holtan-

Hartwig, 2002).

2.4.1.2 Human Alterations to the Denitrification Process and the Consequences

For centuries humans have engaged in activities that influence the nitrogen cycle to benefit

humanity. Activities, such as agriculture, have resulted in an increase in the availability and

movement of nitrogen (Golloway et al., 1995). Such activities have approximately doubled the

rate of nitrogen input into the terrestrial nitrogen cycle, with one of the greatest inputs being

through fertilizer applications. According to Gruber et al., (2008), human activity produces more

than 160 tetragrams (Tg) of new nitrogen per year to be added in terrestrial ecosystems. This

amount surpasses that supplied by BNF on land (110 Tg/yr) or the ocean (140 Tg/yr) (Gruber et

al., 2008). This vast increase in reactive nitrogen contributes to the quality and distribution of

denitrification, enhancing denitrification rates by the addition of nitrate to ecosystems. The

enhancement of denitrification has resulted in numerous problems affecting the earth, one being

the increase in greenhouse gas emissions of nitrogenous species.

Through the mobilization of nitrogen, specifically by denitrification, humans have increased

emissions, transport, reaction and deposition of trace nitrogen gases (nitrous oxide (N2O), nitric

oxide (NO)) into the air (Vitousek et al., 1997). It has been documented that increased fertilizer

inputs is one of the leading causes of greenhouse gas emissions from our agriculture fields

(Dandie et al., 2008; Mosier et al., 2001; Smith et al., 1998). Between 0 and 25% of the applied

nitrogen ends up as N2O and NO gas and denitrification is a major process which contributes to

the emissions of these gases into the atmosphere. Denitrification is responsible for 67% of the

US nitrous oxide emissions (USEPA, 2005), which are increasing at a rate of about 0.2-0.3% per

year (Prinn et al., 1990). Though combustion of fossil fuels is the dominant source of NO

production, global emissions from soils total 5-20 Tg yr-1

(Davidson, 1991).

Page 27: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

17

These anthropogenic changes to the nitrogen cycle are a main driver in regional and global

atmospheric changes. The most significant change seen is the reduction of ozone by N2O.

Nitrous oxide, an un-reactive greenhouse gas in the troposphere, absorbs infrared light, which

contributes to the overall greenhouse warming effect (Albritton et al., 1997). In the stratosphere

N2O photolysis, a chemical reaction in which N2O is broken down by photons, or reacts with

reactive oxygen species, results in the destruction of the ozone (Crutzen and Ehhalt, 1997).

Unlike N2O, NO in the atmosphere is highly reactive and has several effects on atmospheric

reactions.

Nitric oxide comes from destroyed N2O (N2O→N2 + O (1D)) which then reacts with O (

1D)

(N2O + O (1D) → 2NO) (Schlesinger, 1997). NO plays a critical role in affecting both the

concentrations of the hydroxyl radical (OH), a main oxidizing agent in the atmosphere (Logan,

1985), and the photochemical formation of troposphere ozone (O3), a gaseous pollutant (Reich

and Amundson, 1985; Chameides et al., 1994). At high concentrations of atmospheric NO,

oxidation of CO and CH4 leads to the net production of ozone, whereas at low concentrations,

oxidation of CO and CH4 become a sink for ozone (Jacob and Wofsy, 1990). However, the final

oxidation product is nitric acid, a principal component of acid rain.

2.5 References

Albritton DL. (1997) Ozone depletion and global warming. Refrigerants for the 21st Century 1-

5.

Amann R, Ludwig W & Schleifer K. (1995) Phylogenetic identification and in-situ detection of

individual microbial-cells without cultivation. Microbiological reviews 59: 143-169.

Anderson JPE. (1982) SOIL RESPIRATION. Page, A. L.,eds.), pp. P831-872.

Anderson T & Gray T (1990) Soil microbial carbon uptake characteristics in relation to soil-

management. Fems Microbiology Ecology 74: 11-19.

Atlas RM, Horowitz A, Krichevsky M & Bej AK. (1991) Response of microbial-populations to

environmental disturbance. Microbial Ecology 22: 249-256.

Baath E. (1998) Growth rates of bacterial communities in soils at varying pH: A comparison of

the thymidine and leucine incorporation techniques. Microbial Ecology 36: 316-327.

Bacon. (1995) Nitrogen Fertilization in the Environment. pgs 443,446

Page 28: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

18

Baggs EM, Stevenson M, Pihlatie M, Regar A, Cook H & Cadisch G. (2003) Nitrous oxide

emissions following application of residues and fertilizer under zero and conventional

tillage. Plant and Soil 254: 361-370.

Beare M, Coleman D, Crossley D, Hendrix P & Odum E. (1995) A HIERARCHICAL

APPROACH TO EVALUATING THE SIGNIFICANCE OF SOIL BIODIVERSITY TO

BIOGEOCHEMICAL CYCLING. Plant and Soil 170: 5-22.

Beare MH, Coleman DC, Crossley DA, Hendrix PF & Odum EP. (1995) A hierarchical approach

to evaluation the significance of soil biodiversity to biogeochemical cycling. Plant and

Soil 170: 5-22.

Bohannan BJM & Hughes J. (2003) New approaches to analyzing microbial biodiversity data.

Current Opinion in Microbiology 6: 282-287.

Braker G, Schwarz J & Conrad R. (2010) Influence of temperature on the composition and

activity of denitrifying soil communities. Fems Microbiology Ecology 73: 134-148.

Breuer L & Butterbach-Bahl K. (2005) Local temperature optimum of N2O production rates in

tropical rain forest soils of Australia. Australian Journal of Soil Research 43: 689-694.

Burke I, Yonker C, Parton W, Cole C, Flach K & Schimel D. (1989) Texture, climate, and

cultivation effects on soil organic-matter content in US grassland soils. Soil Science

Society of America Journal 53: 800-805.

Calderón FJ, Jackson LE, Scow KM & Rolston DE. (2000) Microbial responses to simulated

tillage in cultivated and uncultivated soils. Soil Biology and Biochemistry 32: 1547-1559.

Cavigelli MA & Robertson GP. (2000) The functional significance of denitrifier community

composition in a terrestrial ecosystem. Ecology 81: 1402-1414.

Chameides WL, Kasibhatla PS, Yienger J & Levy H .(1994) Growth of continental-scale metro-

agro-plexes, regional ozone pollution, and world food-production. Science 264: 74-77.

Chapin FS, III, Walker BH, Hobbs RJ, Hooper DU, Lawton JH, Sala OE & Tilman D. (1997)

Biotic Control over the Functioning of Ecosystems. Science 277: 500-504.

Chaves B, De Neve S, del Carmen Lillo Cabrera M, Boeckx P, Van Cleemput O & Hofman G.

(2005) The effect of mixing organic biological waste materials and high-N crop residues

on the short-time N&lt;sub&gt;2&lt;/sub&gt;O emission from horticultural soil in model

experiments. Biology and Fertility of Soils 41: 411-418.

Cho JC & Tiedje JM. (2000) Biogeography and degree of endemicity of fluorescent

Pseudomonas strains in soil. Applied and Environmental Microbiology 66: 5448-5456.

Page 29: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

19

Conant RT, Paustian K & Elliott ET. (2001) Grassland Management and Conversion into

Grassland: Effects on Soil Carbon. Ecological Applications 11: 343-355.

Crutzen PJ & Ehhalt DH. (1977) Effects of Nitrogen Fertilizers and Combustion on the

Stratospheric Ozone Layer. Ambio 6: 112-117.

Dandie CE, Burton DL, Zebarth BJ, Henderson SL, Trevors JT & Goyer C. (2008) Changes in

bacterial denitrifier community abundance over time in an agricultural field and their

relationship with denitrification activity. Applied and Environmental Microbiology 74:

5997-6005.

Davidson EA & Janssens IA. (2006) Temperature sensitivity of soil carbon decomposition and

feedbacks to climate change. Nature 440: 165-173.

Davidson EA (1991) Fluxes of nitrous oxide and nitric oxide from terrestrial ecosystems.

Enwall K, Nyberg K, Bertilsson S, Cederlund H, Stenstrom J & Hallin S. (2007) Long-term

impact of fertilization on activity and composition of bacterial communities and

metabolic guilds in agricultural soil. Soil Biology & Biochemistry 39: 106-115.

Enwall K, Philippot L & Hallin S. (2005) Activity and composition of the denitrifying bacterial

community respond differently to long-term fertilization. Applied and Environmental

Microbiology 71: 8335-8343.

Enwall K, Throback IN, Stenberg M, Soderstrom M & Hallin S. (2010) Soil Resources Influence

Spatial Patterns of Denitrifying Communities at Scales Compatible with Land

Management. Applied and Environmental Microbiology 76: 2243-2250.

Etchebehere C & Tiedje J. (2005) Presence of Two Different Active nirS Nitrite Reductase

Genes in a Denitrifying Thauera sp. from a High-Nitrate-Removal-Rate Reactor. Appl.

Environ. Microbiol. 71: 5642-5645.

Farquharson R & Baldock J. (2008) Concepts in modelling N2O emissions from land use. Plant

and Soil 309: 147-167.

Ferguson S. (1994) Denitrification and its control. Antonie Van Leeuwenhoek International

Journal of General and Molecular Microbiology 66: 89-110.

Fierer N & Jackson RB. (2006) The diversity and biogeography of soil bacterial communities.

Proceedings of the National Academy of Sciences of the United States of America 103:

626-631.

Fierer N, Schimel J & Holden P. (2003) Influence of drying-rewetting frequency on soil bacterial

community structure. Microbial Ecology 45: 63-71.

Page 30: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

20

Firestone MK, Firestone RB & Tiedje JM. (1980) Nitrous-oxide from soil denitrification-factors

controlling its biological production. Science 208: 749-751.

Franklin RB & Mills AL. (2003) Multi-scale variation in spatial heterogeneity for microbial

community structure in an eastern Virginia agricultural field. Fems Microbiology

Ecology 44: 335-346.

Freckman D, Blackburn T, Brussaard L, Hutchings P, Palmer M & Snelgrove P. (1997) Linking

biodiversity and ecosystem functioning of soils and sediments. Ambio 26: 556-562.

Forestry, Ecology and the Environmental, Vol. 4 (D.Werner aWEN, eds.), pp. 277-342.

Dordrecht, The Netherlands, Springer Netherlands.

Green J & Bohannan BJM. (2006) Spatial scaling of microbial biodiversity. Trends in Ecology &

Evolution 21: 501-507.

Green JL, Bohannan BJM & Whitaker RJ. (2008) Microbial biogeography: From taxonomy to

traits. Science 320: 1039-1043.

Groffman P, Brumme R, Butterbach-Bahl K, Dobbie K, Mosier A & Ojima D. (2000) Evaluating

annual nitrous oxide fluxes at the ecosystem scale. Global biogeochemical cycles 14:

1061-1070.

Gruber N & Galloway JN. (2008) An Earth-system perspective of the global nitrogen cycle.

Nature 451: 293-296.

Hanson G, Groffman P & Gold A. (1994) Denitrification in riparian wetlands receiving high and

low groundwater nitrate inputs. Journal of Environmental Quality 23: 917-922.

Holland E & Coleman D. (1987) Litter placement effects on microbial and organic matter

dynamics in an agroecosystem. Ecology 62: 425-433.

Holtan-Hartwig L, Dorsch P & Bakken LR. (2002) Low temperature control of soil denitrifying

communities: kinetics of N2O production and reduction. Soil Biology & Biochemistry

34: 1797-1806.

Jacob D & Wofsy S (1990) Budgets of reactive nitrogen, hydrocrabonsm and ozone over the

Amazon-forest during the wet season. Journal of geophysical research 95: 16737-16754.

Jones CM, Stres B, Rosenquist M & Hallin S. (2008) Phylogenetic Analysis of Nitrite, Nitric

Oxide, and Nitrous Oxide Respiratory Enzymes Reveal a Complex Evolutionary History

for Denitrification. Molecular Biology and Evolution 25: 1955-1966.

Kandeler E & Murer E. (1993) Aggregate stability and soil microbial processes in a soil with

different cultivation. Geoderma 56: 503-513.

Page 31: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

21

Kennedy A. (1999) Modeling the determinants of species distributions in Antarctica. Arctic,

antarctic, and alpine research 31: 230-241.

Klemedtsson L, Simkins S, Svensson B, Johnsson H & Rosswall T. (1991) Soil denitrification in

3 cropping systems characterized by differences in nitrogen and carbon supply. Water

and NO3 effects on the dentirfication process. Plant and Soil 138: 273-286.

Ladd JN, Foster RC, Nannipieri P & Oades JM. (1996) Soil structure and biological activity.

Marcel Dekker Inc., New York.

Linn D & Doran J. (1984) Effect of water-filled pore-space on carbon-dioxide and nitrous-oxide

production in tilled and nontilled soils. Soil Science Society of America Journal 48:

1267-1272

Logan JA. (1985) Troposheric ozone-seasonal behavior, trands, and anthropogenic influence.

Journal of Geophysical Research-Atmospheres 90: 10463-10482.

Marschner P, Kandeler E & Marschner B. (2003) Structure and function of the soil microbial

community in a long-term fertilizer experiment. Soil Biology and Biochemistry 35: 453-

461.

Martiny JBH, Bohannan BJM, Brown JH, et al., (2006) Microbial biogeography: putting

microorganisms on the map. Nature Reviews Microbiology 4: 102-112.

McCaig AE, Glover LA & Prosser JI. (1999) Molecular analysis of bacterial community

structure and diversity in unimproved and improved upland grass pastures. Applied and

Environmental Microbiology 65: 1721-1730.

Mergel A, Kloos K & Bothe H. (2001) Seasonal fluctuations in the population of denitrifying

and N-2-fixing bacteria in an acid soil of a Norway spruce forest. Plant and Soil 230:

145-160.

Miller MN, Zebarth BJ, Dandie CE, Burton DL, Goyer C & Trevors JT .(2008) Crop residue

influence on denitrification, N2O emissions and denitrifier community abundance in soil.

Soil Biology & Biochemistry 40: 2553-2562.

Mosier AR. (2001) Exchange of gaseous nitrogen compounds between agricultural systems and

the atmosphere. Plant and Soil 228: 17-27.

Nunan N, Wu K, Young IM, Crawford JW & Ritz K. (2002) In Situ Spatial Patterns of Soil

Bacterial Populations, Mapped at Multiple Scales, in an Arable Soil. Microbial Ecology

44: 296-305.

Page 32: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

22

O'Donnell AG, Seasman M, Macrae A, Waite I & Davies JT. (2001) Plants and fertilisers as

drivers of change in microbial community structure and function in soils. Plant and Soil

232: 135-145.

Orians GH. (1997) Biodiversity and terrestrial ecosystem processes. Science Progress 80: 45-63.

Patra AK, Abbadie L, Clays-Josserand A, et al., (2006) Effects of management regime and plant

species on the enzyme activity and genetic structure of N-fixing, denitrifying and

nitrifying bacterial communities in grassland soils. Environmental Microbiology 8: 1005-

1016.

Paul. (2007) Soil Microbiology, Ecology and Biochemistry. 3rd

ed. Elsevier Inc. 2007. 25-53

Paul, E. A. and F. E. Clark (1989). Soil microbiology and biochemistry. San Diego, CA,

Academic Press.

Philippot L, Čuhel J, Saby NP, et al., (2009) Mapping field-scale spatial patterns of size and

activity of the denitrifier community. Environmental Microbiology 11: 1518-1526.

Powlson D, Saffigna P & Kragtcottaar M. (1988) Denitrification at suboptimal temperatures in

soils from different climate zones. Soil Biology & Biochemistry 20: 719-723.

Prosser JI, Bohannan BJM, Curtis TP, et al., (2007) Essay - The role of ecological theory in

microbial ecology. Nature Reviews Microbiology 5: 384-392.

Protection Agency, Washington, DC, April 15, 2005.

Rainey PB, Buckling A, Kassen R & Travisano M. (2000) The emergence and maintenance of

diversity: insights from experimental bacterial populations. Trends in Ecology &

Evolution 15: 243-247.

Reich PB & Amundson RG. (1985). Ambient levels of ozone reduce net photosynthesis in tree

and crop species. Science 230: 566-570.

Rosenfeld JS. (2002) Functional redundancy in ecology and conservation. Oikos 98: 156-162.

Schimel D, Stillwell MA & Woodmansee RG. (1985) Biogeochemistry of C, N, and P in a Soil

Catena of the Shortgrass Steppe. Ecology 66: 276-282.

Schlesinger, W. 1997. Biogeochemistry. Academic Press, Oxford, UK.Sinks: 1990–2003. EPA-

R-05-003. US Environmental

Smith KA, Thomson PE, Clayton H, McTaggart IP & Conen F. (1998) Effects of temperature,

water content and nitrogen fertilisation on emissions of nitrous oxide by soils.

Atmospheric Environment 32: 3301-3309.

Page 33: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

23

Stark JM and Firestone MK Kinetic characteristics of ammonium-oxidizer communities in a

California oak woodland-annual grassland. Soil Biology and Biochemistry 28: 1307-

1317.

Steenwerth KL, Jackson LE, Calderón FJ, Stromberg MR & Scow KM. (2002) Soil microbial

community composition and land use history in cultivated and grassland ecosystems of

coastal California. Soil Biology and Biochemistry 34: 1599-1611.

Stevenson FJ. (1994) Humus chemistry: genesis, composition, reactions. John Wiley and Sons,

New York.

Stotzky G. (1997) Soil as an environment for microbial life. Marcel Dekker Inc., New York.

Stouthamer A. (1991) Metabolic-regulation including Anaerobic metabolism in Paracocus-

denitrificans. Journal of Bioenergetics and Biomembranes 23: 163-185.

Stouthamer AH. (1992) Metabolic Pathways in Paracocus-denitrificans and Closely Related

Bacteria in Relation to the Phylogeny of Prokayotes. Antonie Van Leeuwenhoek

International Journal of General and Molecular Microbiology 61: 1-33.

Swanson FJ, Kratz TK, Caine N & Woodmansee RG. (1988) Landform Effects on Ecosystem

Patterns and Processes. Bioscience 38: 92-98.

Swerts M, Merckx R & Vlassak K. (1996) Influence of carbon availability on the production of

NO, N-2O, N-2 and CO-2 by soil cores during anaerobic incubation (originally published

in Plant and Soil 181: 145-151, 1996). Developments in Plant and Soil Sciences; Progress

in nitrogen cycling studies 633-639.

Szukics U, Abell GCJ, Hödl V, Mitter B, Sessitsch A, Hackl E & Zechmeister-Boltensterns.

Nitrifiers and denitrifiers respond rapidly to changed moisture and increasing temperature

in a pristine forest soil. Fems Microbiology Ecology 72: 395-406.

Torsvik V & Ovreas L. (2002) Microbial diversity and function in soil: from genes to

ecosystems. Current Opinion in Microbiology 5: 240-245.

Torsvik V, Goksoyr J & Daae F. (1990) High diversity in DNA of soil bacteria. Applied and

Environmental Microbiology 56: 782-787.

Torsvik V, Sorheim R & Goksoyr J. (1996) Total bacterial diversity in soil and sediment

communities - A review. Journal of industrial microbiology 17: 170-178.

Travisano M & Rainey PB (2000) Studies of adaptive radiation using model microbial systems.

American Naturalist 156: S35-S44.

USEPA, Inventory of US Greenhouse Gas Emissions

Page 34: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

24

Van Spanning RJ, Delgado, M. J., and Richardson, D. J. (2005) The nitrogen cycle:

Denitrification and its relationship to N2 fixation. Nitrogen Fixation in Agriculture, 281-

320

Vitousek PM, Aber JD, Howarth RW, et al., (1997) Technical Report: Human Alteration of the

Global Nitrogen Cycle: Sources and Consequences. Ecological Applications 7: 737-750.

Wang L, Cai Z, Yang L & Meng L. (2005) Effects of disturbance and glucose addition on nitrous

oxide and carbon dioxide emissions from a paddy soil. Soil and Tillage Research 82:

185-194.

Yachi S & Loreau M. (1999) Biodiversity and ecosystem productivity in a fluctuating

environment: The insurance hypothesis. Proceedings of the National Academy of

Sciences of the United States of America 96: 1463-1468.

Yin B, Crowley D, Sparovek G, De Melo WJ & Borneman J. (2000) Bacterial functional

redundancy along a soil reclamation gradient. Applied and Environmental Microbiology

66: 4361-4365.

Yu Z, Carlson T, Barron E & Schwartz F. (2001) On evaluating the spatial-temporal variation of

soil moisture in the Susquehanna River Basin. Water resources research 37: 1313-1326.

Zumft WG. (1997) Cell biology and molecular basis of denitrification. Microbiology and

Molecular Biology Reviews 61: 533.

Page 35: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

25

CHAPTER 3. OPTIMIZATION OF QPCR ASSAY FOR MICROBIAL DNA

EXTRACTED FROM SOIL

3.1 Introduction

Quantification of gene copy number from DNA extracted from soil microorganisms is a

commonly used and useful tool in microbial ecology (Wallenstein et al., 2006). Currently, the

standard method enabling the measurement of community densities and expression profiles of

community genes involves fluorescent-based quantitative polymerase chain reaction (qPCR).

This method is used because it is sensitive and quantitative. However, the accuracy and precision

of its performance is negatively influenced by inhibitory substances and reaction conditions

(Gallup et al., 2010). Inhibitory substances co-extracted from soil include excess nucleic acids,

humic acids, hemes, polysaccharides, and various proteins (Bustin, 2008; Gallup and Ackerman,

2006). These substances interfere with enzyme kinetics and or reaction chemistries by

overwhelming and/or interfering with enzyme and primer target binding (Gallup et al., 2010)

Various companies have offered different purification kits to minimize inhibitory biological

material carryover (Bustin, 2005, 2008; Bustin et al., 2009; Wilson, 1997) however; they are

expensive and often yield samples that still contain inhibitory polyhenolics and polysaccharides

(Gallup et al., 2010). Because no method is entirely effective at removing inhibitory substances

from all samples, subsequent testing for the presence of inhibition is necessary (Gallup et al.,

2010). The issue of inhibition is further confounded by the fact that different samples, regardless

if they were extracted from the same parent material, have the capability of harboring differing

degrees of inhibitory material (Gallup et al., 2010). Despite the persistence and ubiquity of

inhibition of samples extracted from the environment, researchers have dealt with the issue by

ignoring the problem (Bustin, 2005; Bustin et al., 2009; Nolan and Bustin, 2009). This lack of

resolve has lead to inconstancies within literature (Bustin, 2010). However, methods do exist that

allow for the optimization of qPCR assays on samples contaminated with inhibitory substances.

For example, Gallup and Ackermann proposed the PREXCEL-Q software program to automate

the process of calculating the non-inhibitory dilutions for samples (Gallup et al., 2010). This

method, along with testing reaction conditions such as enzyme mixes, additives, annealing

temperature, and cycling conditions has the potential to improve the reliability of quantification

of gene and gene transcript copy numbers.

Page 36: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

26

To optimize the accuracy and precision of qPCR assays on genomic DNA extracted from soil,

we set out to test the presence of inhibitors and whether reaction chemistries and cycling

conditions influence the efficiencies of qPCR reactions. The aims of this study were to determine

1) the inhibition threshold of genomic DNA samples extracted from soil using PREEXCEL-Q

software (Gallup et al., 2006), 2) whether sampling site and time of sampling influence inhibition

thresholds, 3) whether enzyme chemistries, cycling conditions, and additives influence

sensitivity, repeatability, and qPCR efficiencies. The approach chosen uses a representative

sample, “Stock I”, to test for the dilution range that is unaffected by inhibitory substances but

permits amplification of gene targets. The Stock I approach offers a representative test sample

which serves as both the serially diluted sample for preliminary qPCR test plates (tests for

inhibition range of samples) and as a standard curve material for final qPCR plates for individual

sample/target assessments. Results from these studies will be used to develop a protocol to

streamline optimization of new genes and for quantification of gene transcript copy number for

future use.

3.2 Methods and Results

3.2.1 DNA isolation

Approximately 15 mL of soil was freeze-dried prior to grinding the soil or DNA extraction

(Miller et al., 2008; Griffith et al., 2000; Dandie et al., 2007). Then, approximately 0.25 grams of

soil was placed in a Power bead vial and heated at 70˚C for 5 minutes, vortexed, and reheated

prior to continuing the extraction using the MoBio Power Soil DNA kit (MoBio Laboratories,

Carisbad, CA) per manufacturer instructions. DNA was quantified by adsorption, using a

Nanodrop spectrophotometer.

3.2.2 End point PCR optimization

Prior to experimental applications, primers for 16S rDNA, qnorB, and nosZ (Table 3.1) were

tested via end point PCR to verify amplification of a single target. PCR was carried out with an

Eppendorf Mastercycler Pro (Eppendorf, USA) and Invitrogen Platinum Taq DNA Polymerase

(Invitrogen, Life Technologies, USA), Invitrogen dNTP mix, 5mM MgCl2, 1% Triton X

(Promega Corporation), and primers at 0.4µm (each); and 2 µl of template corresponding to 5 ng

of total DNA were added in 20 µl reaction volumes. Thermal cycling conditions were as follows:

Page 37: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

27

one cycle at 95˚C for 5 min followed by 35 cycles at 95˚C for 30s, annealing temperature (Table

3.1) for 30s, and extension for 72˚C for 30s.

3.2.3 Preparation of stock I

Parameters for the preparation of Stock I were determined using PREXCEL-Q (Gallup et al.,.,.

2005, 2006). In brief, all extracted samples from the Kalsow field site in May were diluted to

create a 1:2 (gDNA: nuclease free water) solution (Table 3.1). Diluted samples were pooled and

nuclease free water was added to make a diluted 1:1 (DNA: water) solution. Stock I was used to

generate an 11-point dilution series (Table 3.2), which was utilized in PREXCEL-Q test plate

(Gallup et al., 2005) to determine inhibition threshold (see section 3.2.4).

3.2.4 Test plate

Quantitative PCRs corresponding to one replicate of an 11-point dilution series of Stock I plus a

no template control (NTC) (Table 3.2) for16S rDNA rDNA, qnorB, and nosZ genes were run to

determine the valid working ranges and inhibitory threshold for samples and standards. In brief,

qPCR was carried out with an Eppendorf Mastercycler ep realplex (Eppendorf, U.S.A) and

SYBR green as the detection system. qPCR for 16S rDNA, qnorB, and nosZ gene number was

performed using SYBR® Green PCR PerfeCTa Master Mix, including AccuStart Taq DNA

Polymerase, dNTP mix, 5 mM MgCl2, SYBR Green I dye, stabilizers, primers (section 3.2.2) at

0.4µm (each), and 6 µl of diluted Stock I were added to 25 µl reaction volumes. Thermal cycling

conditions were as follows: one cycle at 95˚C for 7 min (16S rDNA rDNA and nosZ) and 10 min

(qnorB), followed by 45 cycles at 95˚C for 1s, 58˚C (16S rDNA rDNA and nosZ) and 55˚C

(qnorB) for 20s, and 72˚C for 30s. Fluorescence data collection was measured at 72˚C followed

by a melt curve, and reactions were then held at 4˚C.

To determine inhibition threshold of samples from the Kalsow site, amplification regression

analysis was performed using PREXCEL-Q software. Threshold cycle (Ct) results generated at

log-linear-amplification-capable sample dilutions exhibited a straight line (Bracketed; Figure

3.1), while the inhibitory dilution range generated a curve like a hook at lower sample dilutions

(Circled; Figure 3.1). The hooked area is the dilution range at which samples and reaction are

inhibited; the 1:22 dilution was calculated as the inhibitory threshold for 16S rDNA rDNA,

qnorB, and nosZ gene targets at the Kalsow field site. With the more diluted samples from

Page 38: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

28

sample mixture (Stock I), Cts became log-linear (Figure 3.2), indicating that our sample mixture

behaved without qPCR inhibition when using in well dilutions of 1:140 - 1:3000 (50-500,000

pg/20uL). Further software calculations were performed to generate a dilution (0.008896 ng/uL =

0.009 ng/uL) at which all samples are to be run for study samples.

3.2.5 Experiment 1: Comparison of enzyme performance between Brilliant II, PerfeCTa,

Kappa, and Quantifast

The purpose of the experiment was to determine which enzyme demonstrated the best efficiency

(as measured by slope = -3.21 that corresponds to a perfect doubling of gene copies in each

cycle) and greatest sensitivity (lowest Ct value) with nosZ. The PREXCEL-Q program was used

previously with numerous DNA samples (Stock I; section 3.2.3) to determine the valid working

concentration ranges for all samples (section 3.2.4). The Stock I appeared to behave without

inhibition when used at dilutions of 1:140 - 1:3000. However, because enzyme chemistries vary

in their response to inhibitors, an 11-point dilution series of Stock I (refer to section 3.2.3) was

used with four enzyme kits: Brilliant II (Agilent Technologies), PerfeCTa (Quanta Biosciences,

Inc), SYBR Fast (Kapa Biosystems, Inc.), and QuantiFast (Qiagen). For each enzyme kit,

optimal master mixes and cycling conditions were used as listed in Table 3.3. The annealing

temperature, 60˚C was constant for all enzymes.

Results were analyzed to determine enzyme efficiencies (slopes), relative repeatability (standard

errors) and sensitivity (Ct values). All enzymes exhibited similar threshold inhibition and log-

linear-amplification capabilities (Figure 3.3). Both Brilliant II and Quanta enzyme kits

outperformed Kapa and Quantifast based on efficiencies (slope being closest to -3.21) (Figure

3.4). Qiagen had the best repeatability, based on standard deviation of Ct values, followed by

Brilliant II, Quanta, and Kapa (Table 3.4). When the enzymes were compared for sensitivity, the

ability of an enzyme to amplify the gene target the earliest in a reaction (lowest Ct) the Kapa

enzyme outperformed all others, with the first Ct value occurring at ~ 18 cycles (Figure 3.4).

Due to PCR licensing problems, however, Kapa was not considered for future optimization

work. The Quantifast enzyme did not have good sensitivity or efficiency compared to all other

enzymes so this enzyme kit was also rejected for future optimization work. Because the Brilliant

II enzyme and Quanta enzyme were comparable in efficiency and repeatability, both enzymes

were carried through to Experiment 2.

Page 39: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

29

3.2.6 Experiment 2: Effects of annealing temperature on gene target

The purpose of this experiment was to determine the effect of annealing temperature on the

quantification of gene copy number. Based on the results from the enzyme performance

experiment (Section 2.4), Brilliant II and Quanta were used in this study. Six µl of Sock I serial

dilutions were prepared, and the third (-2) to seventh (-6) dilutions, representing a range of 0.494

ng µl-1

to 0.012 ng µl-1

in well (for calculation see appendix material 3.2.6), were used for DNA

template in reactions. qPCR was performed in replicate 20 µl reactions on 16S rDNA rDNA,

qnorB and nosZ targets following the master mixes and cycling conditions listed in Table 3.5.

Annealing temperatures for cycling were 50 and 60˚C (16S rDNA rDNA and nosZ), and 50 and

55˚C (qnorB) for target molecules.

Results were analyzed based on slopes (efficiencies), repeatability (standard errors) and

sensitivity (low Ct values). For the 16S rDNA rDNA target, the difference in annealing

temperatures did not have an effect on the efficiency, repeatability (Figure 3.5A), or sensitivity

of either enzyme. This lack of difference, specifically in enzyme sensitivity, is likely explained

by its abundance in nature- slight differences in enzyme sensitivity may not be reflected. For

nosZ, both Brilliant and Quanta enzymes resulted in better efficiencies (Figure 3.5B) and

repeatability (lower standard deviations) at higher annealing temperatures (Appendix Table 3.2).

This can be explained by the fact that the annealing temperature was set at the optimal or near

optimal annealing temperatures according to the literature (Braker and Tiedje, 2003; Fierer et al.,

2005; Henry et al., 2006). However, at the higher annealing temperature (60˚C), the sensitivity of

both enzymes was compromised (Figure 3.5B). This is because at higher temperatures the Taq

enzyme is more prudent to which template it binds, ultimately resulting in a lower estimation of

the gene targets copy number. At lower temperatures (Figure 3.5B; Quanta at 50˚C) the Taq

binds to template more readily, resulting in higher gene copy numbers and a better estimation of

the gene abundance in the sample. In comparison, the enzymes were greatly influenced by

annealing temperatures when quantifying the qnorB gene target (Figure 3.5 C). The Quanta

enzyme was more sensitive than Brilliant II enzyme at both annealing temperatures. The

repeatability did not vary significantly between enzymes, but Quanta had the best repeatability of

all the runs at annealing temperature of 55˚C (Appendix Table 3.2). Over all, both enzymes were

comparable in efficiency and repeatability for 16S rDNA rDNA and nosZ. However, Quanta out-

Page 40: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

30

performed Brilliant II in sensitivity for our functional gene targets at annealing temperature of

50˚C for nosZ and at 55˚C for qnorB. Therefore, the Quanta enzyme kit was used in each of the

following experiments.

3.2.7 Experiment 3: Field site and sampling date effect on inhibition threshold

The purpose of this experiment was to determine whether the inhibition of qPCR is influenced

by sampling location (differences in soil type, differences in sampling location) or time of

sampling. Four Stock I stocks (section 3.2.3) were generated to compare qPCR inhibition based

on sampling location and time of sampling. Samples from Kalsow field site and soils from the

Landscape Biomass Project (Iowa State University Uthe Research and Demonstration Farm)

were used to contrast sampling location. Soils sampled from the Kalsow field site in May,

August, and October (Appendix Table 3.3) were used to evaluate the effect of time of sampling

on qPCR inhibition. The enzyme PerfeCTa (Quanta) mix was used in this experiment and all

subsequent experiments based on results from Experiment 2 (section 3.2.6). In brief, an 11-point

Stock I serial dilution, generated from pooled 1:2 diluted DNA (Appendix Table 3.3), + NTC for

each month and site were run in duplicate. qPCR for 16S rDNA, qnorB, and nosZ gene was

performed using SYBR® Green PCR PerfeCTa Master Mix, including AccuStart Taq DNA

Polymerase, dNTP mix, 5 mM MgCl2, SYBR Green I dye, stabilizers, primers (section 3.2.2) at

0.4µm (each), and 4.8 µl of diluted Stock I was added to 25 µl reactions. Thermal cycling

conditions were as follows: one cycle at 95˚C for 10 min, followed by 43 cycles at 95˚C for 1s,

annealing temperature; 50˚C for 20s, and 72˚C for 30s. Fluorescence data collection was

measured at 72˚C, followed by a melt curve, and reactions were then held at 4˚C.

Amplification regression analysis to determine the log-linear amplification ranges (Gallup et al.,

2005), was performed for all Stock I solutions to determine inhibition threshold of samples.

Threshold cycle (Ct) results generated at log-linear-amplification-capable sample dilutions

exhibited a straight line (Bracketed; Figure 3.6), while the inhibitory dilution range generated a

hook-shape curve at lower sample dilutions (Circled; Figure 3.6). This hook area is the dilution

range at which samples and reaction are inhibited (1:22 dilution), which was the same inhibition

threshold for the nosZ gene target at the Kalsow site in May (Figure 3.1; section 3.2.4). With the

more diluted samples (Stock I), Cts were log-linear (Figure 3.7), indicating that our sample

mixture behaved without qPCR inhibition at in well dilutions of 1:140 - 1:3000 (50-500,000

Page 41: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

31

pg/20uL, and sample dilution at 0.008896 ng/uL = 0.009 ng/uL. This was the same dilution

range as the PREXCEL-Q test plate studies (section 3.2.4). Therefore, the most conservative

dilution range for the quantification of gene copy number is from dilution 3 (1:16.7) to dilution 5

(1:109.5) (1 being full concentration) (Table 3.2).

3.2.8 Experiment 4: Effect of annealing temperature on efficiency and sensitivity of enzyme

The purpose of this experiment was to determine the effect of annealing temperature on enzyme

performance. Samples from the Uthe site were used to generate a Stock I (refer to section 3.2.3).

Serial dilutions of the Stock I were made, and dilutions 3, 4, and 5, corresponding to 1:54 –

1:357 in-well dilutions, were used in this assay. The target molecule nosZ was used in this

experiment. qPCR for the nosZ gene was performed using SYBR® Green PCR PerfeCTa

(Quanta) master mix, including AccuStart Taq DNA Polymerase, dNTP mix, 5 mM MgCl2,

SYBR Green I dye, stabilizers, primers (section 3.2.2) at 0.4µm (each), and 4.8 µl of diluted

Stock I were added in 25 µl reaction volumes. Thermal cycling conditions were as follows: one

cycle at 95˚C for 10 min, followed by 43 cycles at 95˚C for 1s, gradient (Table 3.6) at 20s, and

72˚C for 30s. Fluorescence data collection was measured at 72˚C, followed by a melt curve, and

reactions were then held at 4˚C.

Results were analyzed based on slopes (efficiencies) and sensitivity (low Ct values) of the

enzyme at each annealing temperature. As annealing temperature increased (Figure 3.8) the

sensitivity of the enzyme decreased. However, efficiency was more variable (Table 3.7 and 3.8).

As a result of resource constraints, replicates were not conducted. Although the lack of replicates

constrains conclusions that could be drawn from this experiment, values between 55-57˚C gave

the best efficiencies for the PerfeCTa enzyme targeting nosZ.

3.2.9 Experiment 5: Effect of slow cycling with Quanta PerfeCTa SYBR Green SuperMix

The purpose of this experiment was to determine wheter increasing denaturation time would

improve the efficiency and between plate reproducibility of qPCR assays (for background

information refer to Appendix material 3.2.9 A). Kalsow Stock I, dilution range one (full

concentration) through five (1:357 dilution in well), were run in triplicate with two in plate reps

(+2 NTC). qPCR for the nosZ gene was performed using SYBR® Green PCR PerfeCTa

SuperMix Master Mix, including AccuStart Taq DNA Polymerase, dNTP mix, 5 mM MgCl2,

Page 42: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

32

SYBR Green I dye, stabilizers, primers (section 3.2.2) at 0.4µm (each), and 4.8 µl of diluted

Stock I were added in 25 µl reaction volumes. Thermal cycling conditions were as follows: one

cycle at 95˚C for five min, followed by 40 cycles at 95˚C for 30s, 60˚C at 30s, and 72˚C for 30s.

Fluorescence data collection was measured at 72˚C, followed by a melt curve, and reactions were

then held at 4˚C.

Results were analyzed based on slopes (efficiencies) and standard deviation to determine if

repeatability and efficiencies increased when using a Taq (Quanta PerfeCTa SuperMix) designed

for longer denaturation times. Average efficiencies did increase by using PerfeCTa SuperMix

(Table 3.8). Within plate repeatability (Table 3.8) and between plate repeatability (Standard

deviation = 0.18) was also more consistent with longer denaturation time. However, efficiencies

were above 100%, indicating our reaction is not yet optimized (refer to Appendix section 3.2.9

B). Literature searches and phone discussions with Quanta representative Dave Schooster (refer

to Appendix section 3.2.9 B) led us to next test the effects of additives on enzyme efficiency.

3.2.10 Experiment 6: Effect of extended denaturation and additives on qPCR assay

Based on literature searches on additives effect on lowering qPCR reaction efficiencies, the

purpose of this experiment was to test the effect of the additives BSA (400ng µL-1

and 1000ng

µL-1

), T4 gene 32 (10 ng µL-1

and 4 ng µL-1

), and glycerol (5% solution), on the efficiency of

PerfeCTa SYBR Green SuperMix. The goal of this experiment was to see if the addition

additives would lower assay efficiencies below 100% (for background information refers to

Appendix section 2.8 B). Replicate assays were carried out with Kalsow Stock I dilutions two

(1:21.3) through six (1:912.87). qPCR for the nosZ target was performed with master mixes

listed in Table 3.9. Cycling conditions were as follows; one cycle at 95˚C for five mins,

followed by 40 cycles at 95˚C for 30s, 60˚C at 30s and 72˚C for 30s. Fluorescence data

collection was measured at 72˚C, followed by a melt curve, and reactions were then held at 4˚C.

The lower concentrations of additives: BSA (400ng/uL) (Fierer et al., 2006) and T4 gene (80ng)

(Henderson and Dandie, 2010), without glycerol gave reaction efficiencies closest to 100%

(Table 3.10 A and 1.10 B). However, there were no noticeable differences in efficiencies

between runs with BSA and T4 gene added at lower concentrations. Overall, all combinations of

additives lowered efficiencies within a more acceptable range (~ 90%), but the addition of

Page 43: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

33

glycerol had an inconsistent effect on efficiencies. Because of the inconsistencies with the

glycerol addition, it was concluded that glycerol should not be used in optimized reactions.

Additionally, it was concluded that T4 gene should not be used although it worked as well as

BSA. This decision was based on the fact that T4 gene binds to single stranded DNA, including

primers, and can therefore potentially interfere with reactions. Since BSA at lower

concentrations (400ng/uL) gave better efficiencies than reactions with the higher concentration,

BSA at lower concentration was used in optimized reactions.

3.2.11 Experiment 7: Extending denaturation time with Quanta PerfeCTa FastMix and BSA

The purpose of this experiment was to test the effect of a longer denaturation time and the

addition of BSA on the efficiency of Quanta’s FastMix. Based on conversations with Quanta

technicians a denaturation step of 15 seconds was utilized, with the following thermal cycling

conditions: one cycle at 95˚C for 5 mins, followed by 40 cycles at 95˚C for 15s, 60˚C at 20s, and

72˚C for 30s. Fluorescence data collection was measured at 72˚C, followed by a melt curve, and

reactions were then held at 4˚C. BSA at 400ng µL-1

was added to the master mix (see Table 3.9).

Reactions targeting nosZ were run with and without BSA additions (for master mix recipe

without BSA refer to Table 3.5).

The first run failed; diluted Stock I appeared to be degraded after a week at 4˚C and results were

not interpretable. From this result, we concluded that the maximum time Stock I or samples

should be kept in the fridge is 2-3 days. Based on efficiencies of the second run with fresh Stock

I dilutions, reactions without BSA demonstrated efficiencies greater than 100% (Figure 3.9 and

Table 3.11); the addition of BSA lowered the efficiency within the acceptable range of 90-100%.

Therefore, the addition of BSA at a concentration of 400ng is optimal.

3.2.12 Experiment 8: Quanta PerfeCTa SuperMix versus FastMix with BSA

The purpose of this analysis was to determine which enzyme, Quanta SuperMix or Quanta

FastMix, produced a better efficiency. Results from experiment six (section 3.2.10) and

experiment seven (section 3.2.11) were used in this comparative analysis. Overall, there was not

a large difference between enzyme efficiencies (Figure 3.10), although the SuperMix did have

efficiency closer to 100% (0.94) compared to the FastMix (0.91). Both enzyme efficiencies were

Page 44: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

34

in the acceptable range (90-110%) by qPCR standards. Therefore, we determined that either

enzyme could be used for data collection

3.2.13 Experiment 9: qnorB small amplicon (sa) Quanta FastMix with BSA summary

Based on sample results of qnorB long amplicon (637bp) with Quanta SuperMix (data not

shown) it was concluded that the long amplicon was not sufficient for collection of qPCR data.

In brief, efficiencies of replicate runs were below 50%. Literature searches revealed that ideal

amplicon length should be between 100-200bp, but should never exceed 500bp to ensure qPCR

reaction efficiency is as close to 100 % as possible (Optimization guide, Thermo Fisher

Scientific). Therefore, qnorB small amplicon (262bp) primers (Tiedje and Braker, 2003) were

used to determine if acceptable amplification with these primers is possible. Results from qnorB

small amplicon test plate (data not shown), showed that the inhibition threshold was the same as

previous primers tested (refer to section 2.3). Data was collected using Quanta FastMix, long

denaturation with BSA. The Quanta PefeCTa FastMix enzyme was utilized to collect data to

remain consistent with data collection of nosZ and 16S rDNA gene targets.

3.2.14 Experiment 11: nirS test plate with Quanta SuperMix

One replicate of an 11-point dilution series of Stock I and a NTC (Table 3.2.3) were run with

qPCR for nirS functional gene to determine the valid working ranges and inhibitory threshold for

samples and standards. In brief, qPCR was carried out with an Eppendorf Mastercycler ep

realplex (Eppendorf, U.S.A) and SYBR green as the detection system. QPCR for nirS gene target

was performed using SYBR® Green PCR PerfeCTa Master Mix, including AccuStart Taq DNA

Polymerase, dNTP mix, 5 mM MgCl2, SYBR Green I dye, stabilizers, 400ng BSA, primers

(Kandeler et al., 2006) at 0.4µm (each), and 3.8 µl of diluted Stock I were added in 20 µl

reaction volumes. Touchdown thermal cycling conditions, were as follows: one cycle at 95˚C for

2 mins, eight cycles with denaturation at 95˚C for 15s, 63˚C to 56˚C for 30s, extension at 72˚C

for 30s, followed by 32 cycles at 95˚C for 15s, 56˚C for 20s, and 72˚C for 30s. Fluorescence data

collection was measured at 72˚C, followed by a melt curve, and reactions were then held at 4˚C.

Threshold cycle (Ct) results generated at log-linear-amplification-capable sample dilutions

exhibited a straight line (Bracketed; Figure 3.11), while the inhibitory dilution range generated a

hook-like curve at lower sample dilutions (Circled; Figure 3.11). This hook area is the dilution

Page 45: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

35

range at which samples and reaction are inhibited (1:22 dilution), which was the same inhibition

threshold for the nosZ, 16S rDNA rDNA, and qnorB small amplicon gene targets at the Kalsow

site. With the more diluted samples from sample mixture (Stock I), Cts become log-linear

(Figure 3.11, bracketed region), indicating that our sample mixture behaves without qPCR

inhibition at the same in well dilutions (1:140 - 1:3000 (50-500,000 pg/20uL)) and sample

dilution (0.008896 ng/uL = 0.009 ng/uL) as the PREXCEL-Q test plate studies (section 3.2.4).

Therefore, the most conservative dilution range to be utilized in the following studies will be

from three to five (1 being full concentration).

3.2.15 Experiment 12: Comparison of touchdown cycling conditions with nirS gene target

between Quanta SuperMix and Brilliant II

The purpose of this experiment was to determine if there was a difference in efficiency and/or

repeatability using touchdown cycling conditions, a modification of PCR in which the initial

annealing temperature is higher than the optimal temperature of the primers and is gradually

reduced over subsequent cycles until the desired annealing temperature or “touchdown

temperature” is reached thus allowing for a reaction which favors amplification of the desired

amplicon: we used two different enzymes (Quanta SuperMix and Brilliant II). The reason for

revisiting Brilliant II as a possible enzyme was because we had a large surplus of this enzyme: if

there was no difference in efficiency or repeatability between the two enzymes for the nirS gene

then that stock could be used for data collection of this gene. Reactions were run in triplicate. For

Quanta Master Mix refer to section 3.2.14. The SYBR® Green PCR Brilliant II master mix

contained hot start Taq DNA Polymerase, dNTP mix, 5 mM MgCl2, SYBR Green I dye,

stabilizers, 400ng BSA, primers (section 3.2.14) at 0.4µm (each), and 3.8 µl of diluted Stock I

were added in 20 µl reaction volumes. For touchdown thermal cycling conditions for Quanta

refer to section 2.15. Touchdown cycling conditions for Brilliant II are as follows: one cycle at

95˚C for ten min, eight cycles at 95˚C for 30s, 63˚C to 56˚C for 30s, and 72˚C for 60s, followed

by 32 cycles at 95˚C for 30s, 56˚C for 30s, and 72˚C for 60s. Fluorescence data collection was

measured at 72˚C,l followed by a melt curve, and reactions were then held at 4˚C.

The repeatability of within plate runs with Quanta SuperMix was more variable compared to the

repeatability of Brilliant II (Table 3.12). The efficiencies of both enzymes were also variable, but

Brilliant II had better efficiencies compared to the SuperMix (Table 3.12). When Cts were

Page 46: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

36

averaged over the three replicates Brilliant II had an efficiency of 100% compared to the

SuperMix, which had an efficiency of 81%. Overall, the results with the touchdown thermal

cycling conditions were variable when compared to within plate replicates. Although, Brilliant II

did demonstrate better repeatability and greater efficiencies, we next tested the two enzymes with

non-touchdown cycling conditions to see if greater repeatability and efficiencies were produced.

3.2.16 Experiment 13: Comparison of non touchdown cycling conditions with nirS gene target

between Quanta SuperMix and Brilliant II

The purpose of this experiment was to determine whether efficiency and/or repeatability are

affected by non-touchdown cycling conditions using Quanta SuperMix and Brilliant II. For

Quanta master mix refer to section 3.2.14 and for Brilliant II refer to section 3.2.15. Non

touchdown thermal cycling conditions are as follows: one cycle at 95˚C for 2 mins (Quanta) or

10 mins (Brilliant II), followed by 40 cycles at 95˚C for 30s, 56˚C for 30s, and 72˚C for 30s

(Quanta) or 60s (Brilliant II). Fluorescence data collection was measured at 72˚C, followed by a

melt curve, and reactions were then held at 4˚C.

The repeatability of within plate runs with Quanta SuperMix and Brilliant II was comparable, but

Quanta had slightly lower standard errors for Ct values (Table 3.13). Brilliant II demonstrated a

better efficiency compared to Quanta (Table 3.13). When Cts were averaged over two reps

Brilliant II had an efficiency of 107% compared to Quanta, which had an efficiency of 77%.

Overall, the results with the non-touchdown thermal cycling conditions were repeatable when

comparing within plate replicates. Although, Brilliant II did demonstrate better repeatability and

greater efficiencies, we next compared touchdown cycling conditions from section 3.2.15 with

those from this experiment.

3.2.17 Experiment 14: Comparison between touchdown cycling conditions on gene target nirS

with Quanta SuperMix and Brilliant II to no touchdown cycling

This analysis was used to determine whether touchdown or non-touchdown cycling affects

repeatability and efficiency for the nirS gene target. Overall, Brilliant II out performed Quanta

SuperMix when comparing the enzymes’ repeatability and efficiency under both cycling

conditions (Table 3.12 and Table 3.13). When comparing Brilliant II touchdown to Brilliant II

non touchdown reactions, both reaction efficiencies were at ~100% efficiency and had good

within plate repeatability. However, touchdown reactions had smaller standard errors than that of

Page 47: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

37

non-touchdown reactions and better efficiencies. When comparing melt curves (Figure 3.12 A

and B), Brilliant II reaction had a tighter single peak compared to that of non-touchdown

reactions.

3.3 Conclusions

The main objectives of these studies were to determine 1) the inhibitory threshold and working

dilution ranges for gene targets of interest, 2) whether sample location or time of sampling

influenced the inhibitory threshold and working dilution ranges, 3) whether enzyme chemistry

and annealing temperature influenced reaction efficiencies, and 4) optimized reaction conditions

for each gene target. We determined that 1) the inhibitory threshold and working dilution ranges

were the same for all gene targets of interest, 2) sample location and time did not influence

inhibitory threshold and working dilution range, and furthermore 3) enzyme chemistry and

annealing temperatures did influence reaction efficiencies. These results allowed us to optimize

qPCR for our genes of interest and stream-line the process of optimization of future genes.

The mixture inhibitory threshold and working ranges for 16S rDNA rDNA, qnorB small

amplicon, nosZ, and nirS gene targets were the same irrespective of field site location or time of

sampling. Threshold cycle (Ct) results generated at log-linear-amplification-capable sample

dilutions exhibited a straight line for all test plates (Bracketed; Figure 3.2.4, 3.2.5, 3.2.7 and

3.2.14), while the inhibitory dilution range generated a curve like a hook at lower sample

dilutions (Circled; Figure 3.2.4, 3.2.5, 3.2.7 and 3.2.14). The hook areas were the dilution ranges

at which samples and reactions are inhibited (1:22 dilution) either due to humic acids being

coextracted with the DNA or inhibitors introduced elsewhere. The 1:22 dilution, based on

PRECEL-Q software calculations was the same inhibition threshold for all optimized gene

targets. The more diluted samples Cts become log-linear after this threshold, indicating that our

sample mixture behaves without qPCR inhibition at in well dilutions of 1:140 - 1:3000 (50-

500,000 pg/20uL). PRECEL-Q calculations determined our samples optimal dilution at

(0.008896 ng/uL = 0.009 ng/uL) thus allowing room for small discrepancies (pipetting error) to

not interfere with reaction efficiencies. Though, our results have demonstrated that inhibition is

eliminated after the 1:22 dilution we advise that all new genes and samples be run on a test plate

to verify new samples inhibition range.

Page 48: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

38

Enzyme chemistries and cycling conditions greatly influence reaction efficiencies and

sensitivity. Results comparing enzymes (experiments 1 and 12) allowed us to determine which

enzyme resulted in the best efficiencies for the gene targets of interest. By comparing annealing

temperatures (experiments 5 and 7) we were able to determine which annealing temperature

resulted in optimal efficiencies for gene targets. The addition of BSA was determined to be a

standard additive in reactions (experiment 6 and 7), resulting in lower efficiencies in ranges

acceptable for publication. Overall, gene targets nosZ, and qnorB were more greatly influenced

by both enzyme chemistry and annealing temperature than 16S rDNA rDNA gene target (Table

3.4 A). Enzyme chemistry also did not influence gene target nirS (Table 3.12), however, cycling

conditions did (Table 3.12 and 3.13). Based on these results we determined that Quanta

SuperMix would be the standard enzyme for nosZ and qnorB gene targets, although FastMix was

comparable with enzyme repeatability and efficiency. Our preference for Quanta SuperMix was

based on the fact that we were not going to utilize the FastMix for its reduced reaction time

capabilities, meaning our lab has optimized all genes thus far with the longer denaturation and

annealing times. For nirS and 16S rDNA rDNA Brilliant II and Quanta gave comparable results,

but for nirS Brilliant II using touchdown cycling conditions should be utilized. Annealing

temperatures close to published values were also determined to give the best efficiencies for all

gene targets used in these studies.

From these experiments we have developed a protocol to streamline qPCR optimization that

should be used when optimizing new genes.

3.4 References

Braker G & Tiedje JM. (2003) Nitric oxide reductase (norB) genes from pure cultures and

environmental samples. Applied and Environmental Microbiology 69: 3476-3483.

Bustin SA. Real-time quantitative PCR – opportunities and pitfalls. European Pharmaceutical

Review. 2008; 4:18-23. www.europeanphar-qPCR assay development and project

management software www.ijbs.org Int J Biomed Sci vol. 4 no. 4 December 2008 293

maceuticalreview.com

Bustin S. (2005) Real-time, fluorescence-based quantitative PCR: a snapshot of current

procedures and preferences. Expert review of molecular diagnostics 5: 493-498.

Page 49: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

39

Bustin S, Benes V, Garson J, Hellemans J, Huggett J & Kubista M. (2009) The MIQE

Guidelines: Minimum Information for Publication of Quantitative Real-Time PCR

Experiments. Clinical Chemistry 55: 611-622.

Bustin S, Beaulieu J, Huggett J, Jaggi R, Kibenge F & Olsvik P. (2010) MIQE precis: Practical

implementation of minimum standard guidelines for fluorescence-based quantitative real-

time PCR experiments. BMC molecular biology 11: 74.

Dandie CE, Burton DL, Zebarth BJ, Trevors JT & Goyer C. (2007) Analysis of denitrification

genes and comparison of nosZ, cnorB and 16S rDNA rDNA from culturable denitrifying

bacteria in potato cropping systems. Systematic and Applied Microbiology 30: 128-138.

Fierer N, Jackson JA, Vilgalys R & Jackson RB. (2005) Assessment of Soil Microbial

Community Structure by Use of Taxon-Specific Quantitative PCR Assays. Appl.

Environ. Microbiol. 71: 4117-4120.

Fierer N & Jackson RB. (2006) The diversity and biogeography of soil bacterial communities.

Proceedings of the National Academy of Sciences of the United States of America 103:

626-631.

Gallup J & Ackermann M. (2006) Addressing fluorogenic real-time qPCR inhibition using the

novel custom Excel file system 'FocusField2-6GallupqPCRSet-upTool-001' to attain

consistently high fidelity qPCR reactions. Biological procedures online 8: 87-151.

Gallup JM, Sow FB, Van Geelen A & Ackermann MR. (2010) SPUD qPCR Assay Confirms

PREXCEL-Q Software's Ability to Avoid qPCR Inhibition. Current Issues in Molecular

Biology 12: 129-134.

Griffiths BS, Ritz K, Bardgett RD, et al., (2000) Ecosystem response of pasture soil communities

to fumigation-induced microbial diversity reductions: an examination of the biodiversity–

ecosystem function relationship. Oikos 90: 279-294.

Henderson S, Dandie C, Patten C, Zebarth B, Burton D & Trevors J Changes in Denitrifier

Abundance, Denitrification Gene mRNA Levels, Nitrous Oxide Emissions, and

Denitrification in Anoxic Soil Microcosms Amended with Glucose and Plant Residues.

Applied and Environmental Microbiology 76: 2155-2164.

Henry S, Bru D, Stres B, Hallet S & Philippot L. (2006) Quantitative detection of the nosZ gene,

encoding nitrous oxide reductase, and comparison of the abundances of 16S rRNA, narG,

nirK, and nosZ genes in soils. Applied and Environmental Microbiology 72: 5181-5189.

Kandeler E, Deiglmayr K, Tscherko D, Bru D & Philippot L. (2006) Abundance of narG, nirS,

nirK, and nosZ Genes of Denitrifying Bacteria during Primary Successions of a Glacier

Foreland. Applied and Environmental Microbiology 72: 5957-5962.

Page 50: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

40

Miller MN, Zebarth BJ, Dandie CE, Burton DL, Goyer C & Trevors JT. (2008) Crop residue

influence on denitrification, N2O emissions and denitrifier community abundance in soil.

Soil Biology & Biochemistry 40: 2553-2562.

Wallenstein MD, Myrold DD, Firestone M & Voytek M. (2006) Environmental controls on

denitrifying communities and denitrification rates: Insights from molecular methods.

Ecological Applications 16: 2143-2152.

Wilson I. (1997) Inhibition and facilitation of nucleic acid amplification. Applied and

Environmental Microbiology 63: 3741-3751.

Page 51: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

41

Table 3.1: Primers for molecular assays

gene Forward

Primer

Forward

Sequence

Reverse

Primer

Reverse Primer

Sequence

Amplicon

Length bp

Annealing

temperature Source

nosZ nosZ1F

WCSYTGTTCMT

CGACAGCCAG nosZ1R

ATGTCGATCARC

TGVKCRTTYTC 259 60˚C

Henry

et al.,.,.

2006

qnorB qnorB2F GGNCAYCARGG

NTAYGA qnorB7R

ACCCANAGRTGN

ACNACCCACCA 262 51˚C

Braker

&

Tiedje

2003

16S Eub338 ACTCCTACGGG

AGGCAGCAG Eub518

ATTACCGCGGCT

GCTGG 200 55˚C

Fierer

et al.,.,.

2005

Table 3.2: Eleven point dilution series for Stock I solution

Dilution Factor Total Volume

(uL) made Stock I Water

Volume (uL) transferred for

subsequent dilution Dilution 1:

Full concentration 492.6 uL 192.6 uL 300 uL 192.6 uL

10-1 492.6 uL 192.6 uL 300 uL 192.6 uL 2.6

10-2 492.6 uL 192.6 uL 300 uL 192.6 uL 6.5

10-3 492.6 uL 192.6 uL 300 uL 192.6 uL 16.7

10-4 492.6 uL 192.6 uL 300 uL 192.6 uL 42.8

10-5 492.6 uL 192.6 uL 300 uL 192.6 uL 109.5

10-6 492.6 uL 192.6 uL 300 uL 192.6 uL 280.2

10-7 492.6 uL 192.6 uL 300 uL 192.6 uL 716.8

10-8 492.6 uL 192.6 uL 300 uL 192.6 uL 1833.8

10-9 492.6 uL 192.6 uL 300 uL 192.6 uL 4691.0

10-10 492.6 uL 192.6 uL 300 uL 0 uL 12000

Page 52: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

42

Figure 3.1: qPCR inhibitory behavior of Stock I solution at different dilutions. The bracketed Ct

results generated at LOG-linear-amplification-capable sample dilutions. The circled Ct results

are the inhibitory dilution rage at higher template dilutions.

Figure 3.2: qPCR Ct results exhibiting a straight line at LOG-linear-amplification-capable

dilutions

0

5

10

15

20

25

30

35

40

45

-4.5 -4 -3.5 -3 -2.5 -2 -1.5 -1 -0.5 0

Ct

SYB

R

Log (SYBR)

qnorB 16s rDNA nosZ

y = -4.503x + 24.404R² = 0.9955

y = -3.5494x + 10.983R² = 0.9986

y = -3.7971x + 20.356R² = 0.9957

0

5

10

15

20

25

30

35

40

-3 -2.5 -2 -1.5 -1 -0.5 0

Ct

SYB

R

Log (SYBR)

qnorB 16s rDNA nosZ

Page 53: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

43

Table 3.3: Master Mixes and cycling conditions for Experiment 1 - nosZ, Tm = 60C

KAPA Per reaction 40 rxn Final

concentration

Cycling

Conditions

Master Mix 12.5 500 95 3 min

Volume F primer (uL) 2 80 0.4uM 95 3 sec x 45

cycles

Volume R primer (uL) 2 80 0.4uM Tm 20 sec

Water 2.5 100 72 30 sec

Template (uL) 6 Melting curve

Reaction volume 25 1000

QUANTA (FastMix) Per reaction 40 rxn

Master Mix 12.5 500 95 10 min

Volume F primer (uL) 2 80 0.4uM 95 3 sec x 45

cycles

Volume R primer (uL) 2 80 0.4uM Tm 20 sec

Water 2.5 100 72 30 sec

Template (uL) 6 Melting curve

Reaction volume 25 1000

Brilliant II Per reaction 40 rxn 95 10 min

Master Mix 12.5 500 95 30 sec x 45

cycles

Volume F primer (uL) 2 80 0.4uM Tm 30 sec

Volume R primer (uL) 2 80 0.4uM 72 30 sec

Water 2.5 100 Melting curve

Template (uL) 6

Reaction volume 25 1000

Quanitfast Per reaction 40 rxn 95 5 min

Master Mix 12.5 500 95 10 sec x 45

cycles

Volume F primer (uL) 2 80 0.4uM Tm 30 sec

Volume R primer (uL) 2 80 0.4uM 72 30 sec

Water 2.5 100 Melting curve

Template (uL) 6

Reaction volume 25 1000

Page 54: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

44

Figure 3.3: qPCR inhibitory behavior of Stock I solution at different dilutions with enzyme kits:

Brilliant II (diamond), Kappa (square), Quanta (triangle), and Qiagen (x). The bracketed Ct

results generated at LOG-linear-amplification-capable sample dilutions. The hooked Ct results

are the inhibitory dilution rage at higher template dilutions.

Figure 3.4: qPCR Ct results exhibiting a straight line at LOG-linear-amplification-capable

dilutions for enzyme kits Brilliant II, Kapa, Quanta, and Qiagen

20

22

24

26

28

30

32

34

36

38

-4.5 -4 -3.5 -3 -2.5 -2 -1.5 -1 -0.5 0

Ct

log(SYBR)

Brilliant II Kapa Quanta Qiagen

y = -3.1842x + 22.437R² = 0.9999

y = -4.1203x + 17.656R² = 0.9993

y = -3.9856x + 20.362R² = 0.9956

y = -4.1001x + 21.689R² = 0.9954

15

17

19

21

23

25

27

29

31

-2.5 -2 -1.5 -1 -0.5 0

Ct

log(SYBR)

Brilliant II Kapa Quanta Qiagen

Page 55: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

45

Table 3. 4: Experiment 1: Enzyme repeatability performance (n=3)

Enzyme Average Ct Standard deviation Standard error Percent error (%)

Brilliant II 26.3 0.08 0.05 0.001

27.6 0.18 0.10 0.004

28.9 0.17 0.10 0.003

Kapa 22.7 0.32 0.23 0.009

24.3 0.19 0.14 0.006

26.1 0.19 0.14 0.005

Quanta 25.3 0.25 0.14 0.006

26.7 0.35 0.20 0.007

28.5 0.34 0.19 0.007

Qiagen 26.7 0.16 0.09 0.003

28.4 0.15 0.09 0.003

30.1 0.15 0.09 0.003

Table 3.5: Master Mix and Cycling conditions for experiment 2: Effect of annealing temperature

on amplification

Master Mix uL per reaction * 10 reactions Enzyme Cycling conditions

SYBR 10 100 Quanta 95 10 min

PrimerF 1.6 16 95 3 sec x 45

cycles

PrimerR 1.6 16 Tm 20 sec

Water 2 20 72 30 sec

Template 4.8 Melting curve

Brilliant II 95 10 min

95 30 sec x 45

cycles

Tm 30 sec

72 30 sec

Melting curve

Page 56: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

46

y = -3.7588x + 21.978

R² = 0.9829

y = -4.803x + 17.864

R² = 1

y = -3.8457x + 22.018

R² = 0.9922

y = -3.9623x + 21.35

R² = 0.993

0

5

10

15

20

25

30

35

Ct

y = -3.3092x + 11.015

R² = 0.9978

y = -3.6057x + 10.517

R² = 0.9914

y = -3.341x + 11.087

R² = 0.9986

y = -3.281x + 11.384

R² = 0.9901

0

5

10

15

20

25

Ct

Brilliant - 50C Quanta - 50C Brilliant - 60C Quanta-60

Figure 3.5: qPCR Ct results exhibiting a straight line at LOG-linear-amplification-capable

dilutions for enzyme kits Brilliant II (diamond and triangle) and Quanta (square and x) at varying

annealing temperatures for A) 16S rDNA (50˚C and 60˚C), B) nosZ (50˚C and 60˚C), and C)

qnorB (50˚C and 55˚C)

y = -4.6292x + 25.701

R² = 0.9942

y = -5.3205x + 21.077

R² = 0.9973

y = -3.3892x + 30.339

R² = 0.7786

y = -4.4364x + 24.801

R² = 0.9906

0

5

10

15

20

25

30

35

40

45

-3 -2.5 -2 -1.5 -1 -0.5 0

Ct

log(SYBR)

Briliant 50C Quanta 50C Brilliant 55C Quanta 55CC

B

A

Page 57: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

47

Figure 3.6: qPCR inhibitory behavior of different Stock I (Kalsow site in August (diamond) and

October (square) and Uthe site (triangle)) solutions for nosZ gene target at different dilutions

with Quanta enzyme kit. The bracketed Ct results generated at LOG-linear-amplification-

capable sample dilutions. The hooked Ct results are the inhibitory dilution rage at higher

template dilutions.

Figure 3.7: qPCR Ct results exhibiting a straight line at LOG-linear-amplification-capable

dilutions for May, August, and October Kalsow site stock I and Uthe stock I

18

20

22

24

26

28

30

32

34

36

-4.5 -4 -3.5 -3 -2.5 -2 -1.5 -1 -0.5 0

Ct

Log(SYBER)

august october uthe

y = -4.4515x + 18.478R² = 0.9965

y = -3.8668x + 15.997R² = 0.9905

y = -5.5774x + 16.911R² = 0.9301

y = -3.9429x + 17.456R² = 0.9963

15

17

19

21

23

25

27

29

-2.5 -2 -1.5 -1 -0.5 0

Ct

Log(SYBER)

May August October Uthe

Page 58: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

48

Table 3.6: Gradient Conditions for Effect of

Annealing temperature on enzyme performance

Temperature Column

49.8 1

50.1 2

51.0 3

52.3 4

54.0 5

55.8 6

57.7 7

59.5 8

61.2 9

62.6 10

63.6 11

64.1 12

Figure 3.8: qPCR Ct results exhibiting a straight line at LOG-linear-amplification-capable

dilutions stock I at varying annealing temperatures for nosZ gene target

19

21

23

25

27

29

-1.7 -1.5 -1.3 -1.1 -0.9 -0.7

Ct

log (SYBR)

Tm1

Tm2

Tm3

Tm4

Tm5

Tm6

Tm7

Tm8

Tm9

Tm10

Tm11

Page 59: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

49

Table 3.7: Experiment 4: Annealing temperature affect on efficiency of Quanta enzyme

Temperature ™ (˚C) Tm number Regression equation Slope Calculated efficiency

49.8 1 y = -3.8864x + 17.552 3.89 0.81

50.1 2 y = -4.1811x + 17.358 4.18 0.73

51.0 3 y = -3.5311x + 18.25 3.53 0.92

52.3 4 y = -3.8381x + 18.021 3.83 0.82

54.0 5 y = -4.2907x + 17.958 4.29 0.71

55.8 6 y = -3.4822x + 19.363 3.48 0.94

57.7 7 y = -3.4946x + 19.741 3.49 0.93

59.5 8 y = -4.2429x + 19.859 4.24 0.72

61.2 9 y = -5.1249x + 19.39 5.12 0.57

62.6 10 y = -4.6347x + 20.67 4.63 0.64

63.6 11 y = -4.5979x + 21.118 4.59 0.65

64.1 12 N/A (only 2 points)

Table 3 8: Experiment 5: Performance of Quanta PerfeCTa SuperMix enzyme on within plate

repeatability and efficiency

Plate (run) Regression equation Slope Calculated efficiency Standard Deviation

1 y = -3.2061x + 22.142 3.21 1.05 0.07

y = -3.0112x + 22.49 3.01 1.15

2 y = -3.0347x + 22.857 3.03 1.14 0.02

y = -3.0912x + 22.977 3.09 1.11

3 y = -3.1082x + 22.672 3.11 1.10 0.46

y = -2.8315x + 23.032 2.28 1.75

Page 60: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

50

Table 3.9: Experiment 6: Effect of additives Master mixes and cycling conditions Additive 1000ng/uL (Fierer) 400ng/uL ( Kreader) Cycling conditions

BSA Per reaction Per reaction

Quanta Super 10 10 95C 5min

F Primer 1.6 1.6 45 cycles:

R Primer 1.6 1.6 95C 30sec

BSA 2 0.8 60C 30sec

Water 1 2.2 72C 30sec

Template 3.8 3.8 Melt curve

Total 20 20 Hold 4C

BSA+ 5% Glycerol Per reaction Per reaction

Quanta Super 10 10

F Primer 1.6 1.6

R Primer 1.6 1.6

BSA 2 0.8

Water 0 1.2

Glycerol 1 1

Template 3.8 3.8

Total 20 20

200ng/reaction 80ng/reaction

T4 gene Per reaction Per reaction

Quanta Super 10 10

F Primer 1.6 1.6

R Primer 1.6 1.6

BSA 2 0.8

Water 1 2.2

Template 3.8 3.8

Total 20 20

T4 gene + 5%

Glycerol

Per reaction Per reaction

Quanta Super 10 10

F Primer 1.6 1.6

R Primer 1.6 1.6

BSA 2 0.8

Water 0 1.2

Glyceral 1 1

Template 3.8 3.8

Total 20 20

Page 61: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

51

Table 3.10 A: The effect of BSA with and without 5% Glycerol on SuperMix enzyme performance

BSA 1000ng/uL

Name Ct SYBR Amount SYBR

[machfac] Log(SYBR) Efficiency Slope

Play Stock I-3 + BSA

1000ng 24.86 0.153 -0.81531 0.77 -4.0229

Play Stock I-4 +BSA

1000ng 26.24 5.97E-02 -1.22403

Play Stock I-5 + BSA

1000ng 27.81 2.33E-02 -1.63264

Play Stock I-6+ BSA

1000ng 29.81 9.13E-03 -203953

BSA 400 ng/uL

Name Ct SYBR Amount SYBR

[machfac] Log(SYBR) Efficiency Slope

Play Stock I-3 + BSA

400ng 24.92 0.153 -0.81531 0.94 -3.4844

Play Stock I-4 +BSA

400ng 26.37 5.97E-02 -1.22403

Play Stock I-5 + BSA

400ng 27.99 2.33E-02 -1.63264

Play Stock I-6 + BSA

400ng 29.12 9.13E-03 -203953

BSA 1000ng/uL + 5 % glycerol

Name Ct SYBR Amount SYBR

[machfac] Log(SYBR) Efficiency Slope

Play Stock I+ BSA

1000ng 5%Glyc 24.9 0.153

-0.81531 0.89 -3.6261

Play Stock I+ BSA

1000ng 5%Glyc 25.85 5.97E-02

-1.22403

Play Stock I+ BSA

1000ng 5%Glyc 27.69 2.33E-02

-1.63264

Play Stock I+ BSA

1000ng 5%Glyc 29.22 9.13E-03

-203953

BSA 400 ng/uL + 5 % glycerol

Name Ct SYBR Amount SYBR

[machfac] Log(SYBR) Efficiency Slope

Play Stock I-3 + BSA

400ng 5%Glyc 24.92 0.153 -0.81531 0.85 -3.7436

Play Stock I-4 +BSA

400ng 5%Glyc 26.37 5.97E-02 -1.22403

Play Stock I-5 + BSA

400ng 5%Glyc 27.99 2.33E-02 -1.63264

Play Stock I-6 + BSA

400ng 5%Glyc 29.12 9.13E-03 -203953

Page 62: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

52

Table 3.10 B: The effect of T4 gene with and without 5% Glycerol on SuperMix enzyme performance

T4 Gene 200ng

Name Ct SYBR Amount SYBR [machfac] Log(SYBR) Efficiency Slope

Play Stock I-3 + T4 200ng 24.67 0.153 -0.81531 0.90 -3.6016

Play Stock I-4 + T4 200ng 26.11 5.97E-02 -1.22403

Play Stock I-5 + BSA

1000ng 27.46 2.33E-02 -1.63264

Play Stock I-6+ BSA

1000ng 29.12 9.13E-03 -203953

T4 Gene 80ng

Name Ct SYBR Amount SYBR

[machfac] Log(SYBR) Efficiency Slope

Play Stock I-3 + T4 80ng 24.67 0.153 -0.81531 0.94 -3.6016

Play Stock I-4 + T4 80ng 26.11 5.97E-02 -1.22403

Play Stock I-5 + T4 80ng 27.46 2.33E-02 -1.63264

Play Stock I-6 + T4 80ng 29.12 9.13E-03 -203953

T4 Gene 200ng + 5 % glycerol

Name Ct SYBR Amount SYBR [machfac] Log(SYBR) Efficiency Slope

Play Stock I+ T4 200ng

5%Glyc 24.91

0.153 -0.81531 1.10 -3.1485

Play Stock I+ T4 200ng

5%Glyc 26.36

5.97E-02 -1.22403

Play Stock I+ T4 200ng

5%Glyc 27.48

2.33E-02 -1.63264

Play Stock I+ T4 200ng

5%Glyc 28.82

9.13E-03 -203953

T4 Gene 80ng + 5 % glycerol

Name Ct SYBR Amount SYBR

[machfac] Log(SYBR) Efficiency Slope

Play Stock I-3 + T4 80ng

5%Glyc 25.05 0.153 -0.81531 0.99 -3.3297

Play Stock I-4 + T4 80ng

5%Glyc 26.32 5.97E-02 -1.22403

Play Stock I-5 + T4 80ng

5%Glyc 27.67 2.33E-02 -1.63264

Play Stock I-6 + T4 80ng

5%Glyc 29.13 9.13E-03 -203953

Page 63: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

53

Figure 3.9: qPCR Ct results exhibiting a straight line at LOG-linear-amplification-capable

dilutions stock I with extended denaturation time with Quanta PerfeCTa FastMix with A) and

without BSA B) (400ng/ul).

Table 3.11: Experiment 7: Performance of Quanta PerfeCTa FastMix with longer denaturation

time and BSA

Plate (run) Regression equation Slope Calculated efficiency Standard Deviation

1 with BSA y = -3.4884x + 22.178 3.49 0.94 0.03

y = -3.5717x + 22.299 3.57 0.91

y = -3.6623x + 21.952 3.66 0.88

2 without BSA y = -3.0346x + 22.908 3.03 1.14 0.06

y = -3.2381x + 22.405 3.23 1.03

y = -3.2183x + 22.494 2.22 1.05

y = -3.4884x + 22.178

R² = 0.9977

y = -3.5717x + 22.299

R² = 0.9844

y = -3.6623x + 21.952

R² = 0.9872

22

23

24

25

26

27

28

29

30

31

Ct

Rep1 BSA Rep2 BSA Rep3 BSA

y = -3.0346x + 22.908R² = 0.9863

y = -3.2381x + 22.405R² = 0.9909

y = -3.2183x + 22.494R² = 0.9887

22

23

24

25

26

27

28

29

30

-2.5 -2 -1.5 -1 -0.5 0

Ct

Log(SYBR)

Rep1 No BSA Rep2 No BSA Rep3 No BSA

A

B

Page 64: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

54

Figure 3.10: qPCR Ct results exhibiting a straight line at LOG-linear-amplification-capable

dilutions stock I from reactions comparing Quanta PerfeCTa FastMix (diamonds) and SuperMix

(squares) with BSA (400ng)

Figure 3.11: qPCR inhibitory behavior of Stock I solutions for nirS gene target at different

dilutions with Quanta enzyme kit. The bracketed Ct results generated at LOG-linear-

amplification-capable sample dilutions. The hooked Ct results (circled) are the inhibitory dilution

at higher template dilutions.

y = -3.5741x + 22.143

R² = 0.9915

y = -3.4865x + 22.179R² = 0.9975

24.00

25.00

26.00

27.00

28.00

29.00

30.00

-2.5 -2 -1.5 -1 -0.5 0

CT

LOG (SYBR)

FastMix SuperMix

0

5

10

15

20

25

30

35

40

-4.5 -4 -3.5 -3 -2.5 -2 -1.5 -1 -0.5 0

Ct

LOG (SYBR)

nirs

Page 65: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

55

Table 3.12: Experiment 12: Performance of Quanta and Brilliant II with longer denaturation and

touchdown thermal cycling conditions

Quanta SuperMix

Rep 1 Ct SYBR Slope Efficiency Average Ct Standard

deviation

Standard

error Average slope

25.09 -3.3392 0.99 24.49 0.69 0.49 0.81

26.87

26.20 0.83 0.59

27.56

27.52 0.33 0.23

29.4

29.33 0.28 0.20

Rep 2 24.11 -4.528 0.66

25.7

27.1

29.8

Rep 3 24.26 -3.8066 0.83

26.04

27.92

28.81

Stratagene Brilliant II

Ct SYBR Slope Efficiency Average Ct

Standard

deviation

Standard

error Average slope

Rep 1 24.19 -3.3165 1.00 23.93 0.23 0.13 1.00

25.05

24.92 0.20 0.11

26.64

26.35 0.33 0.19

28.17

28.23 0.15 0.09

Rep 2 23.86 -3.7651 0.84

24.69

26.43

28.4

Rep 3 23.74 -3.4543 0.94

25.03

25.98

28.12

Page 66: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

56

Table 3.13: Experiment 13: Performance of Quanta and Brilliant II with longer

denaturation and no touchdown thermal cycling conditions

Brilliant II Avg CT Std Dev Std Error Slope Efficiency

23.63 0.61 0.43 -3.155 1.07

24.51 0.64 0.45

25.61 0.09 0.07

27.56 0.22 0.16

Quanta 23.77 0.05 0.04 -4.0274 0.77

24.99 0.37 0.26

26.81 0.25 0.18

28.64 0.05 0.04

Figure 3.12: Melt curves of nirS amplicons with Brilliant II enzyme run with non touchdown

(A) and touchdown (b) cycling conditions

A.

B.

Temperature [°C]

9492908886848280787674727068666462

- dI / dT

(%

)

100

90

80

70

60

50

40

30

20

10

0

-10

-20

-30

-40

Threshold: 33%

Temperature [°C]

959493929190898887868584838281807978777675747372717069686766656463626160

- dI / dT

(%

)

100

90

80

70

60

50

40

30

20

10

0

-10

-20

-30

-40

Threshold: 33%

Page 67: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

57

CHAPTER 4. SPATIO-TEMPORAL DISTRIBUTION OF QNORB AND NOSZ-

BEARING DENITRIFIERS IN AN AGRO-ECOSYSTEM AND A PRAIRIE

ECOSYSTEM

4.1 Abstract

Knowing the distribution of denitrifying microorganisms may increase our understanding of the

relationship between microbial ecology and ecosystem processes. Using molecular analyses, we

explored the spatio-temporal distribution of denitrifiers in relation to soil properties at two scales,

landform and land management, over a growing season. Soil samples were taken in May,

August, and October and 13 soil physiochemical properties in concert with abundance of norB

and nosZ-bearing denitrifying bacteria were measured. Neither the absolute or relative

abundances of denitrifiers to total bacteria were influenced by land management. However,

denitrifier abundance was strongly influenced by soil type and sampling date. Our results suggest

that the abundance of denitrifier genes is dependent on soil properties intrinsic to soil/ landform

position rather than land management. Of the soil properties measured, nitrogen species and iron

and copper consistently explained the pattern observed in nosZ gene abundance through time and

space. By contrast, qnorB abundance did not correlate with physiochemical properties measured

in this study, but instead positively correlated with nosZ abundance. Overall our findings suggest

that the variation detected in qnorB and nosZ abundance is dependent on physiological responses

of the organism to cellular cues (e.g. micronutrients or O2 consintrations), reflecting the ability of

microbial communities to respond to resource changes in their environment specifically N,

copper and iron.

4.2 Introduction

Ecosystem processes, such as denitrification, exhibit high spatial and temporal variability

(Nunan et al., 2002), and are therefore difficult to predict. Because microorganisms mediate

many ecosystem processes, knowing the factors that influence their distribution may help explain

the variation observed in process rates. The distribution of microorganisms within a soil matrix is

influenced by the soil’s inherent heterogeneity, which is a result of chemical (i.e. pH, carbon (C),

nitrogen (N)), physical (e.g. soil structure), and biological (i.e. root systems) characteristics that

vary over both space and time. As a result, understanding the influence of soil properties on the

distribution of microorganisms is an important first step to understanding the ecological impact

Page 68: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

58

of these organisms.

One useful approach for characterizing the distribution of microorganisms in their environment

is biogeography, which can offer insight into the environmental and historical pressures shaping

the community of interest (Cho and Tiedje, 2000; Bohannan and Huges, 2003; Fierer and

Jackson, 2006). Biogeography studies designed with natural environmental gradients allow for

comparative analysis that facilitate our understanding of environmental factors influencing

microbial community structure over time and space (Martiny et al., 2006). In addition to this

approach, functional-trait-based biogeography studies may provide insight into the distribution

of microbial functional guilds of interest. With the advancement of environmental molecular

biology, novel tools for identifying patterns of microbial functional guilds have been developed;

yet very little is still known about the distribution of functional genes (Philippot et al., 2009). By

utilizing DNA-based techniques, studies targeting genes in the environment provide a tractable

method for characterizing spatial patterns of microbial functional guilds (Enwall et al., 2010;

McGill et al., 2006; Philippot et al., 2009).

Due to their significant role in nitrogen cycling, including the production and reduction of the

greenhouse gas nitrous oxide (N2O), denitrifying bacteria have become a popular model

community for the study of microbial ecology (Enwall et al., 2005; Philippot and Hallin, 2005).

Previous studies have examined denitrifier abundance, composition, and diversity by amplifying

bacterial genes involved in denitrification, including, nitrate reductase (napA and narG), nitrite

reductase (nirS and nirK), nitric oxide reductase (qnorB and cnorB) and nitric oxide reductase

(nosZ). While many investigations have focused on analysis of edaphic factors and process rates

in relation to denitrifier composition (Cuhel et al., 2010; Enwall et al., 2010; Enwall and Hallin,

2009; Hallin et al., 2009; Ma et al., 2008; Rich et al., 2004), the biogeography of the

corresponding genes (nir-type, norB, and nosZ) has been limited or ignored. This has been

especially true with respect to qnorB and cnorB, the genes that encode nitric oxide reductase, the

enzyme directly responsible for the production of the greenhouse gas nitrous oxide (N2O).

Because norB and nosZ encode for the enzymes responsible for the production and reduction of

N2O, changes in the population density of norB- and nosZ-bearing denitrifiers may affect N2O

Page 69: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

59

emissions. Three recent studies have looked at how the abundance of norB-type and nosZ genes

relate to soil physiochemical properties (Dandie et al., 2007; Dandie et al., 2008; Miller et al.,

2008); however, these studies either found weak correlations or did not find significant

correlations or between denitrifier community densities and measured soil properties. For

example Dandie et al. (2008) demonstrated variation in gene abundance of cnorB, nosZ, and

nirK in both space and time, but these changes were not related to any environmental parameter

measured (NO3-, NH4

+, EOC, pH, WFPS); they did detect weak relationships were between

cnorB copy number, EOC, and temperature. Miller et al (2008) also did not observe an effect of

C (glucose) or NO3- additions on total bacterial and nosZ copy numbers, but did observe an

increase in cnorB copy numbers in microcosm experiments. The generality of these results

remain unclear because the cnorB primers were designed to target denitrifiers specific to the

study site, and additionally, microcosm studies require validation at field scale. Therefore field

studies trying to understand the distribution of the genes responsible for N2O production and

reduction are warranted.

The environmental parameters known to influence denitrifier abundance, denitrification rates,

and gas fluxes at the field scale (i.e. C, N, water-filled pore space (WFPS), and pH) (Wallenstein

et al., 2006) are influenced by numerous factors of biogeography, including: soil

physiochemistry, topography, land management, and sampling date. The availability of water

and nutrients is directly controlled by interactions between land management, landform, and

season (Franzluebbers et al., 2000; Kassen and Rainey, 2004; Sabiene et al., 2010). Although

land management and landform changes are distal controls of denitrification (Wallenstein et al.,

2006), land management mainly influences both nutrient availability (crop cover and nitrogen

additions) and soil disturbance (tillage) (Bruns et al., 1999; Stres et al., 2004), whereas landform

(topography and location of soil series) and soil type affect water and nutrient distribution (Yates

et al., 2006a; Yates et al., 2006b). Additionally, seasonal variation influences the interaction of

soil structure, water content, nutrient movement, and nutrient availability. Together, these factors

create a multidimensional habitat, which has been shown to influence denitrification rates and

may also alter the distribution of norB and nosZ gene abundance over time and space.

Page 70: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

60

Although important to denitrifying enzymes, trace elements such as copper (Cu) and iron (Fe)

are not commonly tested in field experiments. Both nitric (Nor) and nitrous (Nos) oxide

reducates are metalloezymes, requiring metal cofactors, Fe (Nor) and Cu (Nos) respectively, for

activation (Dreusch et al 1996). In addition to their importance in enzyme activity, these metal

ions also elicit a regulatory response by directly affecting the biosynthesis of the denitrification

components dependent on these metals (Haltia et al., 2003; Labbe et al., 2003; Zumft et al.,

1997). For example, bacterium switching to denitrying conditions requires Fe for the de novo

synthesis of Fe-containing enzymes and cytochrome electron carriers (Hendrik et al., 1998;

Zumft et al., 1997). Additionally, Cu has been demonstrated to regulate the outer membrane

protein of NosA (Hubbar et al., 1989). Therefore, enzymatic cofactors may be critical factors

regulating the microbial genetic potential, particularly when considering genes encoding for

metaloenzymes such as those found in the denitrification pathway.

Knowing the relationship between the distribution of norB and nosZ genes and the spatial and

temporal variations of soil properties may reveal what soil mechanisms maintain the genetic

potential for denitrification. This is a critical first step for linking denitrifying organisms to

process level measurements, which may ultimately provide insight to the variation detected into

N2O vs. N2 production (Hallin et al., 2009; Stres et al., 2008). The main objectives of this study

were to determine the extent to which norB and nosZ gene abundances vary spatially and

temporally, and to determine the environmental factors associated with landform, land

management, and sampling date that explain the variation in gene abundance. We hypothesize

that 1) the abundance of norB and nosZ will be greatest in low-lying sites, where nutrient

availability (C and N) and water content are greatest; 2) abundance of norB and nosZ will be

greater in the prairie due to greater nutrient availability (EOC) and wetter conditions; 3)

fluctuations in the abundances of norB and nosZ over time will correlate with changes in soil

chemistry; and 4) the abundance of norB and nosZ will be influenced by the concentrations of

enzymatic co-factors associated with nitric oxide reductase and nitrous oxide reductase. To test

these hypotheses, the abundances of norB and nosZ were quantified from four soil series along

soil catena, under two management regimes, at three time points over a growing season. In

addition, abundance of the total bacterial community (16S rRNA) was determined to quantify

changes in the relative abundance of the denitrification genes.

Page 71: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

61

4.3 Materials and Methods

4.3.1 Experimental site

Soil samples were collected from two sites, Kalsow Prairie, a 160-acre tall grass prairie state

preserve located in southern Pocahontas County, Iowa, U.S.A (42° 34´N, 94° 34´ W), and an

adjacent field south of the prairie under cultivated corn-corn-soybean rotation. The prairie was

never plowed, but was used as a hay field and pasture until 1948, whereas the agriculture field

has been managed following conventional practices. Both sites are located on the Des Moines

Lobe Landform, which is part of the prairie pothole region, and was last glaciated approximately

13,000 years ago (Prior, 1991). Annual precipitation for the county over the past 57 years was

30.34 in, with an average of 8.84, 12.62, 6.42, and 2.46 inches of precipitation in spring (March-

May), summer (June-August), fall (September-November), and winter (December-February),

respectively. The annual mean temperature ranges from -8.8˚C in January to 22.7˚C in July

(Carlson and Todey, 2009). Each site has the same gently sloping (~10%) topography and soil

series: Clarion (summit), Webster (side-slope), Canisteo (foot-slope), and Okoboji (closed

depression) (Soil survey staff). The taxonomic class of the Clarion series is a fine loam, mixed,

superactive, mesic Typic Hapludolls, Webster and Canisteo series are a fine, clay loam, mixed,

superactive, mesic Typic Endoaquolls, and Okoboji series is a fine silty clay loam, smectitic,

mesic Cumulic Vertic Endoaquoll.

4.3.2 Soil sampling

Three transects containing the four soil types were established in both the prairie and agriculture

field, and within each transect a 2 m2 plot was established for each soil series. In May, August,

and October of 2009 seven soil cores (2.2 cm in diameter and 10 cm in depth) were collected

from each plot. The cores were composited and homogenized, and then divided into two

subsamples, one for physiochemical analysis that was stored at 4˚C and the other for nucleic acid

analysis that was stored at -80˚C.

4.3.3 Physiochemical analysis

Soil inorganic nitrogen (NO3--N and NH4

+-N) was measured colorimetrically from 2M KCl

extracts, and summed to derive total inorganic nitrogen (TIN). Extractable organic carbon (EOC)

Page 72: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

62

was analyzed from diluted 2M KCl extracts on a TOC analyzer (Elementar liquiTOC, Americas

Inc., Mt. Laurel, NJ, USA). Total nitrogen (TN) and carbon (TC) was measured via combustion

(LECO Instruments CNS 600) (Bremer, 1996; Nelson and Sommers; 1996). Soil pH was

measured using a pH meter (Accumet & Basic AB15 pH meter, Fisher Scientific) in a 1:2 soil:

water (vol:vol) slurry. Iron and copper availability were determined using the Melich-3 test at the

Iowa State University Plant and Soil Analysis Laboratory. Soil moisture was determined

gravimetrically by oven-drying subsamples of field moist soil for 24 hours at 105˚F. Gravimetric

weight and bulk density measurements

(http://websoilsurvey.nrcs.usda.gov/app/WebSoilSurvey.aspx) were then used to determine the

water-filled pores space (WFPS) of each soil type.

4.3.4 DNA Extraction

Soil was freeze-dried the day prior to grinding (Miller et al., 2008; Klammer et al., 2000; Griffith

et al., 2000; Dandie et al., 2007). DNA was extracted from ground soil using the MoBio

PowerSoil DNA kit (MoBio Laboratories, Carisbad, CA) with the following modifications. The

Power bead vial containing 0.25g soil was heated for 5 min at 70˚C, vortexed, and reheated prior

to continuing the extraction using the manufacturer’s recommended procedure. DNA was

quantified by adsorption, using a Nanodrop spectrophotometer.

4.3.5 qPCR of norB and nosZ abundance

Quantitative polymerase chain reaction (qPCR) was used to quantify the abundance of the total

bacterial community (16S rRNA) and denitrifiers possessing nitric oxide reductase (qnorB and

cnorB) and nitrous oxide reductase (nosZ) genes using previously described primers (Fierer et

al., 2005; Braker and Tiedje, 2003; Henry et al., 2006). Although cnorB (Braker and Tiedje,

2003), the homolog of qnorB, is known to be present in a wide array of denitrifiers, we were

unable to satisfactorily amplify this gene with previously published primer sets, which precluded

their inclusion in this study.

Reactions were carried out in an Eppendorf Mastercycler ep realplex (Eppendorf) using SYBR®

Green PCR PerfeCTa (Quanta Biosciences, Inc. Gaithersburg, MD USA) Master mix, which

included AccuStart Taq DNA Polymerase, dNTP mix, MgCl2, SYBR Green I dye, and the

Page 73: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

63

stabilizer dimethyl sulfoxide. Additionally each reaction contained 0.4 µm of each primer, 400ng

BSA (New England Biolabs, Inc), and 2.8 ng of total DNA in 20 µl reaction volumes. BSA was

used to enhance qPCR efficiency. Thermal cycling conditions were as follows: 1 cycle at 95˚C

for 5 min followed by 41 cycles at 95˚C for 30s, 60˚C (16S rRNA and nosZ) or 56˚C (qnorB) for

30s, and 72˚C for 30s. Fluorescence was collected during extension at 72˚C. Dissociation curves

were performed for all wells, and the results confirmed the specificity of the PCR product

formation. Two independent qPCR reactions were performed on each soil sample. Two no-

template controls were run in each assay and no-template controls did not generate product or

produced negligible values. Standard curves were generated via Stock I method (Gallup and

Ackermann 2008) and compared to plasmid standard curves generated from serial dilutions of

PCR amplified clone gene inserts containing 16S rRNA, qnorB, and nosZ genes to calculate gene

copy number. Determination of the inhibition-free DNA concentration used in these reactions

was determined using the PREXCEL-Q method (Gallup and Ackermann 2008).

4.3.6 Statistical analysis

Individual soil physical and chemical properties and gene abundances were first analyzed using

descriptive statistical methods and data were inspected for departures from normality using

residual plots and the Shapiro-Wilk test and data were transformed when needed to meet

normality assumptions. The relative abundance of denitrifiers was based on calculating the ratios

of the qnorB and nosZ genes to total bacteria (16S rRNA). To evaluate the extent to which the

absolute and relative abundance of qnorB, nosZ, and 16S rRNA genes differed among land

management regimes and soil type over the growing season, we performed a repeated measures

analysis (PROC MIXED; SAS Inst. 2003) with time, land-use, and soil-type as the main effects.

We tested differences among means using Tukey’s honestly significant difference. Possible

interacting relationships between soil properties and absolute or relative gene abundances were

modeled with all-possible-subset regression, and variables were chosen when Akaike’s

Information Criterion was a minimum (Littell et al., 1996). Refer to Supplementary Table 1 for

the list of the 15 variables include in statistical analysis. The results were considered statistically

significance at P value of < 0.05 and marginally significant P < 0.10.

Page 74: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

64

4.4 Results

The abundance of bacteria estimated by using qPCR of the 16S rRNA gene ranged from 8.0 x107

to 2.0 x1011

copies g-1

dry soil, and averaged 1.3 x1010

copies g-1

dry soil over the three sampling

dates. As expected, abundances of nitric (qnorB) and nitrous (nosZ) oxide reductase-bearing

bacteria were less than the estimated total bacteria. The abundance of qnorB ranged from 2.7

x103 to 9.5 x10

5 copies g

-1 dry soil with an average of 2.5 x10

4 copies g

-1 dry soil, while the

abundance of nosZ ranged from 4.9 x105 to 1.0 x10

8 copies g

-1 dry soil, averaging 5.7 x10

6

copies g-1

dry soil. The abundance of both denitrifier genes was within the range reported in

previous studies that quantified the abundance of the homolog gene cnorB, and nosZ (Dandie et

al., 2007, Dandie et al., 2008; Henry et al., 2006; Miller et al., 2009; Philippot et al., 2009). The

proportion of the total bacterial community possessing qnorB or nosZ was approximately

0.001% and 0.06%, respectively, which is similar to values previously reported in other

temperate soils (Dandie et al., 2008; Henry et al., 2006; Bru et al., 2010). Similar to other studies

(Ma et al., 2008) we did not detect an effect of land management on that the absolute and relative

abundance of denitrifier genes. However, denitrifier abundance was strongly influenced by soil

type and sampling date.

4.4.1 Soil series: Averaging among management regimes, nosZ abundance differed among soil

series with the greatest abundance in Canisteo compared to Webster or Okoboji soils (F3,12=

5.45, P < 0.012; Figure 4.1). Clarion had 3.8% greater nosZ abundance compared to Webster

soil. Differences in nosZ abundance among soil series could be explained in part by aboitic and

biotic properties that varied with soil series (Supplementary Table 4.2). For example, soil pH,

EOC and available C: N ratio were significantly greater in the foot-slope soil (Canisteo), whereas

iron, copper, and NO3- concentrations were significantly greater in the closed depression soil

(Okoboji), compared to other soils in the catena. Interestingly, WFPS, NH4, and TN

concentrations were not significantly different among the different soil types.

The abundance of qnorB and the micronutrients copper and iron were the best predictors of nosZ

abundance (Table 4.1). For example, qnorB abundance in conjunction with WFPS and

concentrations of C, N, and metal cofactors explained 77-78% of the variation in nosZ

abundance in the Webster and Okoboji soils (Table 4.1). In Clarion soil, 44% of nosZ variation

Page 75: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

65

was explained by TN and Cu concentrations, whereas soil pH, C: N ratio, and Fe best explained

the variation observed in nosZ in Canisteo soil, producing a combined r2 of 0.70.

4.4.2 Sampling date: Patterns in 16S rRNA, qnorB, and nosZ abundance due to sampling date

(Table 2) were apparent in combined data sets from both management regimes, which revealed a

greater than 2-fold, transient increase in gene abundance in August. Peak gene abundances

occurring in August were evident at both a land management (16S rRNA and nosZ; Figure 4.2 A

and B) and landform scale (nosZ; Figure 4.2 C). For example, in Canisteo and Okoboji soils,

nosZ abundance increased 10% from May to August and decreased 10 and 13%, respectively,

from August to October (P≤ 0.03, Figure 4.2 C), whereas in Webster soil, a 13% decrease in

abundance was observed from August to October (P≤ 0.002).

Parallel to seasonal differences in gene abundance, there were significant seasonal increases in

TN (+17%), and Cu (+8%) concentrations (P ≤ 0.02) and a 20% increase in EOC (P = 0.06)

from May to August, followed by significant decreases in TN (-19%), EOC (-19%), and Cu (-

10%) from August to October (P ≤ 0.03; Figure 4.3 A, B, C). In contrast, WFPS and C: N ratio

demonstrated an inverse relationship with all three gene abundances, significantly decreasing by

15 and 18% from May to August and significantly increasing by 23 and 21% from August to

October (P ≤ 0.02; Figure 4.3 D, E). Because fluctuations of multiple physiochemical properties

could concurrently affect denitrifiers gene abundance, interaction effects were modeled by all

subsets regression. In May, soil pH, TN, Fe, and qnorB abundance explained 47% of the

variation observed in nosZ abundance. In August, EOC and TN explained 30% of the variation

observed in nosZ abundance, whereas in October, TN, TIN concentrations and qnorB abundance

best explained the variation, producing a combined r2 of 0.52 (Table 4.3). In only October, 38%

of the variation observed in qnorB abundance was explained by nosZ abundance.

4.5 Discussion

The objectives of this study were to determine the extent to which qnorB and nosZ gene

abundances vary spatially and temporally, and to determine the environmental factors associated

with landform, land management, and sampling date that predicted these differences. We found

that 1) nosZ abundance varied by landform, but the greatest abundances were not only found in

Page 76: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

66

low-lying soils as we predicted; 2) land management practices were not observed to influence the

abundance of qnorB or nosZ; 3) temporal variations were observed in qnorB and nosZ

abundance, with N, Cu, and Fe, predicting gene abundance in a time-dependent manner; and 4)

micronutrients Cu and Fe were important in predicting the abundance of nosZ-bearing

denitrifiers along the catena.

Our study demonstrates that the landform and soil series within land management better

described the denitrifier community abundance than land management regime. Although studies

have observed variation in denitrifier abundance at the land management scale (Enwall et al.,

2005; Philippot et al., 2009), our study diverges from previous studies that demonstrated

significant differences in denitrifier nir-type and nosZ gene abundances between N-fertilizer and

non-fertilized lands (Chéneby et al., 2009; Enwall et al., 2005; Hallin et al., 2009). Instead, our

results may provide support for the genetic resilience of these organisms to perturbations (Wertz

et al., 2007); considering the diversity of soil denitrifiers, this functional group maybe

sufficiently dissimilar to maintain their stability in numbers despite the differences in land

management at our site. Another explanation for the lack of difference in gene abundance

between land managements is that changes in N inputs due to fertilizer events are ephemeral and

inherent soil characteristics (more stable forms of N and C) are more important to explaining

denitrifier abundance. Several studies have demonstrated that soil type (i.e. characterization

based on physical structure) does influence microbial community size and structure (Grivan et

al., 2003; Ma et al., 2008). For example, Girvan et al., (2003) found that five fields containing

Stagnogley soil on two geographically distinct farms exhibited near identical microbial

composition profiles despite varying chemical characteristics, such as organic matter and N

concentrations. Ma et al., (2008) likewise observed that nosZ abundance was not influenced by

land management (cultivated vs. non cultivated prairie potholes) but by landform (position

within the potholes) in Saskatchewan, Canada. The agricultural and prairie field in our study

contain potholes with the same soil series profile (Clarion, Webster, Canisteo, and Okoboji),

suggesting that the significant difference in denitrifier gene abundance among soil series (Figure

4.1) are dependent upon the local underlying soil chemistry and structure.

Page 77: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

67

The greater abundance of nosZ in Canisteo (toe-slope) and Clarion (summit) soils compared to

Webster and Okoboji soils suggest that resources and environmental conditions intrinsic to these

soils favor higher nosZ-bearing denitrifier population sizes. The variability in nosZ abundance

was not predicted by uniform suite of soil properties along the catena (Table 4.1). However N

(TN and C: N) and the metal cofactors (copper or iron) consistently correlated with nosZ

abundance (Table 4.1), highlighting the importance of these elements to nosZ-bearing

denitrifiers. TN is the pool of soil N able to be mineralized and oxidized into substrates used as

the terminal electron acceptors in denitrification; copper acts as a cofactor for the Nos enzyme

(Coyle et al., 2005; Brown et al., 2000), and iron acts as a cofactor for the Nor enzyme, and is

needed by cytochrome for electron transfer (Labbe et al., 2003). Previous studies have shown

that differences in physiochemical characteristics (e.g soil pH, organic matter, WFPS) contribute

to spatially distinct microbial communities (Carelli et al., 2000; Cavigelli et al., 1995; Gorlenko

et al., 1997; Kreitz et al., 1997; Marschner et al., 2007). Although not measured in this study,

differences in denitrifier microbial community composition may also be contributing to the

variation in the suite of response variables associated with denitrifier gene abundance along the

catena (Table 4.1). The denitrifier community possessing nosZ is made up of an assemblage of

many taxa which can grow under both aerobic and anaerobic conditions. Therefore, the diversity

of the soil conditions driving the distribution of nosZ-bearing denitrifiers may reflect the

taxonomic diversity of this functional group.

Seasonal variation in resource availability (Grayston et al., 2001; Zak et al., 2003) and

temperature (Grayston et al., 2001; Rache et al., 2010) within the soil habitat may have fostered a

shift in the total bacterial and denitrifier microbial community abundances. Here, differences in

16S rRNA, qnorB, and nosZ abundances coincided with shifts in edaphic factors (Figure 4.3 A,

B, C, D, E) and changes in ambient air temperature, suggesting that seasonal changes in soil

physiochemical properties and/or temperature drive shifts in microbial abundance (Bardgett et

al., 1999; Bell et al., 2008; Berg and Rosswall, 1987; Braker et al., 2010; Desnues et al., 2007).

At peak plant biomass production (August), when the warmest temperatures and the greatest

abundances of denitrifiers were observed, export of labile carbon, organic acids, and amino acids

in root exudates have been shown to be greatest (Zak et al., 2003). This flush of available

nutrients has been shown to stimulate mineralization and increase denitrifier gene abundance

Page 78: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

68

(Baudoin et al., 2009; Enwall et al., 2010; Dandie et al., 2008; Li et al., 1995) and activity

(Bremer et al., 2009) in other studies. Our results support this idea, with EOC and TN peaking in

August (Figure 4.3) concurrent with maximum abundances of 16S rRNA, qnorB, and nosZ genes.

Moreover, this difference in gene abundance was also observed for nosZ at both land scales

(Figure 4.2B and Figure 4.2C), with multiple resources (i.e. N, Fe, EOC) predicting this shift

throughout the growing season (Table 4.3). Together these data suggest that seasonal fluctuations

in the genetic potential for denitrification may be tightly liked to belowground plant inputs

(Bremer et al., 2007; Henry et al., 2008). Even so, it is interesting to note that we were unable to

detect a response in gene abundance due to differences in belowground inputs of perennial native

prairie versus continuous corn (land management scale).

Although rarely considered in studies linking microbial ecology to ecosystem processes, a

novelty of our study was to determine the importance of enzymatic cofactors, specifically copper

and iron, to denitrifier abundance. Indeed, for nosZ abundance, both trace metals were significant

predictors of nosZ distribution both spatially and temporally. This correlation may demonstrate

the importance of these enzymatic cofactors in the biological productivity of nosZ-bearing

microorganisms (Matsubara et al., 1982; Labbe et al., 2003; Zumft et al., 1997) in the field.

Several studies have demonstrated that the addition of Cu and Fe to growth medium of

denitrifiers causes denitrification rates to increase, resulting in decreases in N2O and increases in

N2 emissions, while removing these trace metals demonstrates an inverse reaction with increased

N2O production (Bruland et al., 1991; Grange and Ward, 2002; Matsubara et al., 1982). Because

N2O binds very poorly to metal ions and is difficult to activate towards its reduction, the

evolution of a multi-copper (12 Cu ions) active center capable of catalyzing the reduction of N2O

to N2 gas was evolved (Iwasaki et al., 1980; Granger and Ward, 2003). In addition, Fe is needed

by cytochrome for electron transfer (Labbe et al., 2003), FNR (fumarate nitrate reduction

regulatory protein; regulates transcription of denitrification genes in response to O2

concentrations), and is a metal cofactor for Nor. Our results indicate that enzymatic cofactors

may be crucial for regulating microbial genetic potential, particularly when considering genes

encoding for enzymes in the denitrification pathway.

Page 79: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

69

Unlike that of nosZ, the abundance of qnorB was not observed to correlate to environmental

parameters measured in this study, but did correlate to nosZ abundance. Although most studies

use nir-type genes to correspond to N2O production, and find relationships between these genes

and edaphic factors (Enwall et al., 2010; Philippot et al., 2009), our results suggest that the

distribution of qnorB (the gene directly responsible for encoding the enzyme that produces N2O

gas) may not be sensitive to or regulated by the soil properties measured in this study. Similarly,

Dandie et al. (2007) found only weak relationships between EOC, temperature and cnorB

abundance. Additionally, studies using mutational analysis have shown that the transcription of

nor genes are regulated by FNR-type protein, a transcriptional regulator activated under O2

limited or anoxic conditions (De Boer et al., 1996). Thus the distribution and regulation of this

gene may instead depend on the physiological response of the organisms to its surrounding

environment, such as physiological cues of oxygen concentrations or NO production (Van

Spanning et al., 2005; Zumft et al., 2002).

4.6 Conclusion

By characterizing the spatio-temporal distribution of total bacterial, qnorB-bearing and nosZ-

bearing denitrifier abundances at the landform and land management scales, over three different

sampling dates, our study demonstrated that 1) the abundance of denitrifier genes can be

dependent on soil properties intrinsic to landform position rather than land management, and 2)

seasonal changes in denitrifier abundance reflect the ability of microbial communities to respond

to resource changes in their environment, with TN being the most consistent explanatory

variable. Despite heterogeneity among soil series and time, our study indicates that the

abundance of nosZ is consistently related to TN, Cu and Fe, while qnorB is correlated with nosZ

abundance rather than edaphic characteristics. Our results consistently support the idea that

plant-microbe interactions can illicit seasonal variations in the genetic potential for

denitrification. In addition, we found that soil concentrations of enzymatic cofactors Fe and Cu

strongly influence nosZ abundance. Future studies investigating the regulation of denitrifier

communities should include micronutrients, particularly those studies investigating the ratio of

N2O: N2 production. Because nir-type gene abundance has been correlated with denitrification

activity (Cuhel et al., 2010; Graham et al., 2010; Mosier and Francis, 2010), it is possible that

differences in nosZ-bearing abundances in response to changes in environmental parameters over

Page 80: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

70

time (McGill et al., 2010) may contribute to differences in N2O vs. N2 emissions at field scale

(Phillipot et al., 2010). However, microorganisms possessing these genes belong to a diverse

group and some of the populations involved in denitrification may not be active or contribute

marginally to denitrification rates at any given point in time (Attard et al., 2010). Therefore, to

further increase our level of understanding about the biogeographic distribution of denitrifiers,

future studies will require the incorporation of phylogenetics and proteomic analysis, along with

approaches utilizing mRNA to better aid our understanding of how individual denitrifiers

respond to their environment.

Acknowledgments

This work was supported by the Department of Ecology, Evolution and Organismal Biology,

Iowa State University, Ames, IA. We would like to thank John McCleary from Iowa’s DNR,

Adam Pintar for statistical advice; and the many people who have helped us in the field: Steve

Adams; Matthew Raymond; and Stith Wiggs. In addition we would like to thank Steve Adams,

Elizabeth Bach, and Ryan Williams for insightful feedback to improve this manuscript.

4.7 References

Attard, E., S. Recous, et al. (2010). "Soil environmental conditions rather than denitrifier

abundance and diversity drive potential denitrification after changes in land uses." Global

Change Biology 17(5): 1975-1989.

Bardgett RD, Leemans DK, Cook R & Hobbs PJ (1997) Seasonality of the soil biota of grazed

and ungrazed hill grasslands. Soil Biology and Biochemistry 29: 1285-1294.

Baudoin E, Philippot L, Chèneby D, et al. (2009) Direct seeding mulch-based cropping increases

both the activity and the abundance of denitrifier communities in a tropical soil. Soil

Biology and Biochemistry 41: 1703-1709.

Bell C, McIntyre N, Cox S, Tissue D & Zak J (2008) Soil microbial responses to temporal

variations of moisture and temperature in a Chihuahuan Desert Grassland. Microbial

Ecology 56: 153-167.

Page 81: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

71

Berg P & Rosswall T (1987) Seasonal variations in abundance and activity of nitrifiers in four

arable cropping systems. Microbial Ecology 13: 75-87-87.

Bohannan BJM & Hughes J (2003) New approaches to analyzing microbial biodiversity data.

Current Opinion in Microbiology 6: 282-287.

Braker G, Schwarz J & Conrad R (2010) Influence of temperature on the composition and

activity of denitrifying soil communities. Fems Microbiology Ecology 73: 134-148.

Bremer C, Braker G, Matthies D, Reuter A, Engels C & Conrad R (2007) Impact of plant

functional group, plant species, and sampling time on the composition of nirK-Type

denitrifier communities in soil. Applied and Environmental Microbiology 73: 6876-6884.

Brown K, Tegoni M, Prudencio M, et al. (2000) A novel type of catalytic copper cluster in

nitrous oxide reductase. Nat Struct Mol Biol 7: 191-195.

Bru D, Ramette A, Saby NPA, et al. (2010) Determinants of the distribution of nitrogen-cycling

microbial communities at the landscape scale. ISME J.

Bruns MA, Stephen JR, Kowalchuk GA, Prosser JI & Paul EA (1999) Comparative Diversity of

Ammonia Oxidizer 16S rRNA Gene Sequences in Native, Tilled, and Successional Soils.

Appl. Environ. Microbiol. 65: 2994-3000.

Carelli M, Gnocchi S, Fancelli S, Mengoni A, Paffetti D, Scotti C & Bazzicalupo M (2000)

Genetic Diversity and Dynamics of Sinorhizobium meliloti Populations Nodulating

Different Alfalfa Cultivars in Italian Soils. Appl. Environ. Microbiol. 66: 4785-4789.

Cavigelli M, Robertson G & Klug M (1995) Fatty-acid Methyl-ester (FAME) Profiles as

Measure of Soil Microbial Community Structure. Plant and Soil 170: 99-113.

Page 82: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

72

Cavigelli MA & Robertson GP (2000) The functional significance of denitrifier community

composition in a terrestrial ecosystem. Ecology 81: 1402-1414.

Cavigelli MA & Robertson GP (2001) Role of denitrifier diversity in rates of nitrous oxide

consumption in a terrestrial ecosystem. Soil Biology & Biochemistry 33: 297-310.

Cheneby D, Brauman A, Rabary B & Philippot L (2009) Differential Responses of Nitrate

Reducer Community Size, Structure, and Activity to Tillage Systems. Applied and

Environmental Microbiology 75: 3180-3186.

Cho JC & Tiedje JM (2000) Biogeography and degree of endemicity of fluorescent

Pseudomonas strains in soil. Applied and Environmental Microbiology 66: 5448-5456.

Coyle CL, Zumft WG, Kroneck PMH, KÖRner H & Jakob W (1985) Nitrous oxide reductase

from denitrifying. European Journal of Biochemistry 153: 459-467.

Cuhel J, Simek M, Laughlin RJ, Bru D, Cheneby D, Watson CJ & Philippot L (2010) Insights

into the Effect of Soil pH on N2O and N-2 Emissions and Denitrifier Community Size

and Activity. Applied and Environmental Microbiology 76: 1870-1878.

Dandie CE, Burton DL, Zebarth BJ, Trevors JT & Goyer C (2007) Analysis of denitrification

genes and comparison of nosZ, cnorB and 16S rDNA from culturable denitrifying

bacteria in potato cropping systems. Systematic and Applied Microbiology 30: 128-138.

Dandie CE, Burton DL, Zebarth BJ, Henderson SL, Trevors JT & Goyer C (2008) Changes in

bacterial denitrifier community abundance over time in an agricultural field and their

relationship with denitrification activity. Applied and Environmental Microbiology 74:

5997-6005.

De Boer, A. P. N., J. van der Oost, et al. (1996). Mutational analysis of the <em>nor</em> gene

cluster which encodes nitric-oxide reductase from Paracoccus denitrificans. European

Journal of Biochemistry, Wiley-Blackwell. 242: 592-600.

Page 83: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

73

De Graaff M-A, Classen AT, Castro HF & Schadt CW (2010) Labile soil carbon inputs mediate

the soil microbial community composition and plant residue decomposition rates. New

Phytologist 188: 1055-1064.

Deiglmayr K, Philippot L, Hartwig UA & Kandeler E (2004) Structure and activity of the

nitrate-reducing community in the rhizosphere of Lolium perenne and Trifolium repens

under long-term elevated atmospheric pCO2. Fems Microbiology Ecology 49: 445-454.

Desnues C, Michotey VD, Wieland A, Zhizang C, Fourçans A, Duran R & Bonin PC (2007)

Seasonal and diel distributions of denitrifying and bacterial communities in a hypersaline

microbial mat (Camargue, France). Water Research 41: 3407-3419.

Enwall K & Hallin S (2009) Comparison of T-RFLP and DGGE techniques to assess denitrifier

community composition in soil. Letters in Applied Microbiology 48: 145-148.

Enwall K, Philippot L & Hallin S (2005) Activity and composition of the denitrifying bacterial

community respond differently to long-term fertilization. Applied and Environmental

Microbiology 71: 8335-8343.

Enwall K, Throback IN, Stenberg M, Soderstrom M & Hallin S (2010) Soil Resources Influence

Spatial Patterns of Denitrifying Communities at Scales Compatible with Land

Management. Applied and Environmental Microbiology 76: 2243-2250.

Feng Y, Motta AC, Reeves DW, Burmester CH, van Santen E & Osborne JA (2003) Soil

microbial communities under conventional-till and no-till continuous cotton systems. Soil

Biology & Biochemistry 35: 1693-1703.

Fierer N & Jackson RB (2006) The diversity and biogeography of soil bacterial communities.

Proceedings of the National Academy of Sciences of the United States of America 103:

626-631.

Page 84: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

74

Florinsky IV, McMahon S & Burton DL (2004) Topographic control of soil microbial activity: a

case study of denitrifiers. Geoderma 119: 33-53.

Franzluebbers AJ, Hons FM & Zuberer DA (1995) Tillage-induced seasonal changes in soil

physical properties affecting soil CO2 evolution under intensive cropping. Soil and

Tillage Research 34: 41-60.

Girvan MS, Bullimore J, Pretty JN, Osborn AM & Ball AS (2003) Soil type is the primary

determinant of the composition of the total and active bacterial communities in arable

soils. Applied and Environmental Microbiology 69: 1800-1809.

Gorlenko MV, Majorova, T.N., Kozhevin, P.A (1997) Disturbances and their influence on

substrate utilization patterns in soil microbial communities. Microbial Communities 84-

93.

Graham DW, Trippett C, Dodds WK, et al. (2010) Correlations between in situ denitrification

activity and nir-gene abundances in pristine and impacted prairie streams. Environmental

Pollution 158: 3225-3229.

Granger, J. and B. B. Ward (2003). "Accumulation of Nitrogen Oxides in Copper-Limited

Cultures of Denitrifying Bacteria." Limnology and Oceanography 48(1): 313-318.

Grayston SJ, Griffith GS, Mawdsley JL, Campbell CD & Bardgett RD (2001) Accounting for

variability in soil microbial communities of temperate upland grassland ecosystems. Soil

Biology and Biochemistry 33: 533-551.

Green JL, Bohannan BJM & Whitaker RJ (2008) Microbial biogeography: From taxonomy to

traits. Science 320: 1039-1043.

Page 85: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

75

Hallin S, Jones CM, Schloter M & Philippot L (2009) Relationship between N-cycling

communities and ecosystem functioning in a 50-year-old fertilization experiment. Isme

Journal 3: 597-605.

Henderson S, Dandie C, Patten C, Zebarth B, Burton D & Trevors J (2010) Changes in

Denitrifier Abundance, Denitrification Gene mRNA Levels, Nitrous Oxide Emissions,

and Denitrification in Anoxic Soil Microcosms Amended with Glucose and Plant

Residues. Applied and Environmental Microbiology 76: 2155-2164.

Henry S, Bru D, Stres B, Hallet S & Philippot L (2006) Quantitative detection of the nosZ gene,

encoding nitrous oxide reductase, and comparison of the abundances of 16S rRNA, narG,

nirK, and nosZ genes in soils. Applied and Environmental Microbiology 72: 5181-5189.

Kassen R & Rainey PB (2004) The ecology and genetics of microbial diversity. Annual Review

of Microbiology 58: 207-231.

Kennedy NM, Gleeson DE, Connolly J & Clipson NJW (2005) Seasonal and management

influences on bacterial community structure in an upland grassland soil. Fems

Microbiology Ecology 53: 329-337.

Kong AYY, Hristova K, Scow KM & Six J (2010) Impacts of different N management regimes

on nitrifier and denitrifier communities and N cycling in soil microenvironments. Soil

Biology and Biochemistry 42: 1523-1533.

Kreitz S & Anderson TH (1997) Substrate utilization patterns of extractable and non-extractable

bacterial fractions in neutral and acidic beech forest soils. Microbial Communities 149-

160.

Labbe, N., S. Parent, et al. (2003). "Addition of trace metals increases denitrification rate in

closed marine systems." Water Research 37(4): 914-920

Littell RC, Milliken GA, Stroup WW, Wolfinger RD (1996) SAS system for mixed models. SAS

Publishing, Cary, North Carolina, USA

Page 86: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

76

Li Z, Li L, Pan Y & Wu S (1995) Influence of the root exudates of wheat shoots on the

denitrifiers in rhizosphere. Turang Xuebao 32: 408-413.

Ma WK, Bedard-Haughn A, Siciliano SD & Farrell RE (2008) Relationship between nitrifier and

denitrifier community composition and abundance in predicting nitrous oxide emissions

from ephemeral wetland soils. Soil Biology and Biochemistry 40: 1114-1123.

Magalhaes C, Costa J, Teixeira C & Bordalo AA (2007) Impact of trace metals on denitrification

in estuarine sediments of the Douro River estuary, Portugal. Marine Chemistry 107: 332-

341.

Marschner P & Rengel Z (2007) Nutrient cycling in terrestrial ecosystems. Springer-Verlag

GmbH, Heidelberg Germany.

Martiny JBH, Bohannan BJM, Brown JH, et al. (2006) Microbial biogeography: putting

microorganisms on the map. Nature Reviews Microbiology 4: 102-112.

Matsubara, T., K. Frunzke, et al. (1982). "Modulation by copper of the products of nitrite

respiration in Pseudomonas perfectomarinus." J. Bacteriol. 149(3): 816-823.

McGill BJ, Enquist BJ, Weiher E & Westoby M (2006) Rebuilding community ecology from

functional traits. eds.), pp. 178-185. Portland, OR.

McGill BM, Sutton-Grier AE & Wright JP (2010) Plant Trait Diversity Buffers Variability in

Denitrification Potential over Changes in Season and Soil Conditions. PLoS ONE 5: 8.

Mergel A, Kloos K & Bothe H (2001) Seasonal fluctuations in the population of denitrifying and

N-2-fixing bacteria in an acid soil of a Norway spruce forest. Plant and Soil 230: 145-

160.

Page 87: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

77

Miller MN, Zebarth BJ, Dandie CE, Burton DL, Goyer C & Trevors JT (2008) Crop residue

influence on denitrification, N2O emissions and denitrifier community abundance in soil.

Soil Biology & Biochemistry 40: 2553-2562.

Mosier AC & Francis CA (2010) Denitrifier abundance and activity across the San Francisco

Bay estuary. Environmental Microbiology Reports 2: 667-676.

Nunan N, Wu K, Young IM, Crawford JW & Ritz K (2002) In Situ Spatial Patterns of Soil

Bacterial Populations, Mapped at Multiple Scales, in an Arable Soil. Microbial Ecology

44: 296-305.

Philippot L & Hallin S (2005) Finding the missing link between diversity and activity using

denitrifying bacteria as a model functional community. Current Opinion in Microbiology

8: 234-239.

Philippot L, Čuhel J, Saby NP, et al. (2009) Mapping field-scale spatial patterns of size and

activity of the denitrifier community. Environmental Microbiology 11: 1518-1526.

Pintathong, P., D. Richardson, et al. (2009). "Influence of metal ions and organic carbons on

denitrification activity of the halotolerant bacterium, Paracoccus pantotrophus P16 a

strain from shrimp pond." Electronic Journal of Biotechnology 12(2).

Rich JJ & Myrold DD (2004) Community composition and activities of denitrifying bacteria

from adjacent agricultural soil, riparian soil, and creek sediment in Oregon, USA. Soil

Biology & Biochemistry 36: 1431-1441.

Sabiene N, Kusliene G & Zaleckas E (2010) The influence of land use on soil organic carbon

and nitrogen content and redox potential. Zemdirbyste-Agriculture 97: 15-24.

Soil Survey Staff, Natural Resources Conservation Service, United States Department of

Agriculture. Web Soil Survey. Available online at http://websoilsurvey.nrcs.usda.gov/

accessed 05/29/2009

Page 88: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

78

Stres B, Mahne I, Avgustin G & Tiedje JM (2004) Nitrous oxide reductase (nosZ) gene

fragments differ between native and cultivated Michigan soils. Applied and

Environmental Microbiology 70: 301-309.

Stres B, Danevcic T, Pal L, et al. (2008) Influence of temperature and soil water content on

bacterial, archaeal and denitrifying microbial communities in drained fen grassland soil

microcosms. Fems Microbiology Ecology 66: 110-122.

Wallenstein MD, Myrold DD, Firestone M & Voytek M (2006) Environmental controls on

denitrifying communities and denitrification rates: Insights from molecular methods.

Ecological Applications 16: 2143-2152.

Wertz S, Dandie CE, Goyer C, Trevors JT & Patten CL (2009) Diversity of nirK Denitrifying

Genes and Transcripts in an Agricultural Soil. Appl. Environ. Microbiol. AEM.01588-

01509.

Wolsing M & Prieme A (2004) Observation of high seasonal variation in community structure of

denitrifying bacteria in arable soil receiving artificial fertilizer and cattle manure by

determining T-RFLP of nir gene fragments. Fems Microbiology Ecology 48: 261-271.

Yates, T. T, SI, et al. (2006) Wavelet spectra of nitrous oxide emission from hummocky terrain

during spring snowmelt. Soil Science Society of America, Madison, WI, ETATS-UNIS.

Yates TT, Si BC, Farrell RE & Pennock DJ (2006) Probability distribution and spatial

dependence of nitrous oxide emission: Temporal change in hummocky terrain. Soil

Science Society of America Journal 70: 753-762.

Zak DR, Holmes WE, White DC, Peacock AD & Tilman D (2003) Plant diversity, soil microbial

communities, and ecosystem function: Are there any links? Ecology 84: 2042-2050.

Page 89: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

79

The model r2 is the cumulative r

2from the model

The P value listed is the model P value from repeated measures ANOVA from PROC MIX

analysis.

Table 4.1 All subsets regression results testing for the

ability of edaphic and biological factors to predict nosZ abundance

Model

variables

Parameter

Model r2

P

TN Clarion 0.44 0.022

Cu 0.013

TC Webster 0.77 0.097

TN 0.024

qnorB 0.0001

Soil pH Canisteo 0.70 0.025

Fe

C:N

0.004

0.002

Iron Okoboji 0.78 0.062

WFPS 0.050

Cu 0.053

TC 0.013

TN 0.007

qnorB 0.005

Table 4.2 Abundances of 16S rRNA, qnorB, and nosZ genes at each sampling date.

Significant differences indicated by different letters (P < 0.05).

May August October P

16S rRNA (x 1010

copies) 10.1A 10.7

B 10.1

A <0.0001

qnorB (x 104 copies) 4.5

A 4.8

B 4.3

A 0.006

nosZ (x 107 copies) 6.8

A 7.3

B 6.6

A <0.0001

Page 90: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

80

The model r2 is the cumulative r

2from the additive effects of model variables

Table 3 All subsets regression in identifying edaphic and biological factors predicting

nosZ and qnorB abundance by sampling date

Dependent

variable

Model

variable

Parameter

Model r2

P

nosZ pH

Iron

qnorB

TN

May 0.47 0.066

0.016

0.014

0.003

TN

EOC

August 0.30 0.049

0.013

TIN

TN

qnorB

October 0.52 0.012

0.011

0.0006

qnorB -- May -- N.S

-- August -- N.S

nosZ October 0.38 0.002

Page 91: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

81

Figure 4.1: nosZ abundance averaged for each soil type. Bars with the same letter were not

significantly different (P ≤ 0.05). Error bars represent ± standard error.

Page 92: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

82

Figure 4.2: Sampling date effects on (A & B) 16S rRNA and nosZ by land management and C)

on nosZ by landform. Bars accompanied by the same letter between months were not

significantly different (ANOVA; α ≤ 0.05). Error bars represent ± standard error.

Page 93: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

83

Figure 4.3: Sampling date effects on A) Cu , B) EOC, C) TN, D) WFPS, E) C:N Means

accompanied by the same letter were not significantly different (ANOVA; α ≤ 0.05). Error bars

represent ± standard error.

Page 94: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

84

APPENDIX DATA CHAPTER 3

Appendix Table 3.1: Kalsow field site DNA concentrations from May sampling for generation of

Stock I solution

Stock I - Play DNA Used for Experiments 1 & 2 Quantified 2.16.2010

Transect Soil Series 260 280 260/280 260/230 ng/uL ng/uL

K1 Cn 0.418 0.186 2.24 2.86 20.9 20.9

Ct 0.535 0.296 1.81 2.14 26.7 26.7

Ok 0.393 0.221 1.78 2.1 19.7 19.7

W 0.459 0.219 2.1 2.74 23 23

K2 Cn

Ct 0.347 0.167 2.08 2.29 17.4 17.4

Ok 0.226 0.121 1.86 2.47 11.3 11.3

W 0.01

0.32 0.12 1 1

K2 Cn 0.494 0.236 2.1 0.92 24.7 24.7

Ct 0.327 0.138 2.36 3.88 16.3 16.3

Ok 1.127 0.791 1.42 1.29 56.3 28.15

W 0.259 0.114 2.26 2.16 13 13

C1 Cn 0.294 0.124 2.37 3.29 14.7 14.7

Ct 0.167 0.059 2.82 0.1 8.3 8.3

Ok 0.055 0.003 16.55 7 2.7 2.7

W 0.046 0.046 1.01 2.92 2.3 2.3

C2 Cn 0.482 0.277 1.74 0.76 24.1 24.1

Ct 0.117 0.058 2 2.67 5.8 5.8

Ok 0.135 0.082 1.6 2.95 6.8 6.8

W 0.103 0.048 2.13 3.24 5.1 5.1

C3 Cn 0.556 0.399 1.42 0.87 28.3 28.3

Ct 0.356 0.188 1.9 0.65 17.8 17.8

Ok 0.125 0.085 1.48 1.24 6.3 6.3

W 0.315 0.199 1.58 0.61 15.8 15.8

14.79 9.28913043

Note: K2Ok was pre-diluted 1:2 before adding to Stock I

Page 95: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

85

APPENDIX DATA CHAPTER 3 CONTINUED

Appendix Table 3.2: Experiment 2: enzyme repeatability performance (n=2)

Enzyme Gene Target Annealing

Temperature

Average Ct Standard

Deviation

Standard

error

Brilliant II 16S rDNA 50 13.8 0.06 0.04

15.1 0.18 0.13

16.3 0.07 0.05

17.7 0.08 0.06

19.2 0.08 0.06

Quanta 13.7 0.03 0.02

14.7 0.08 0.06

16.2 0.03 0.02

17.8 0.20 0.15

19.5 0.06 0.06

Brilliant II 16S rDNA 60 13.9 0.02 0.02

15.2 0.04 0.03

16.5 0.05 0.04

17.8 0.11 0.08

19.4 0.16 0.11

Quanta 60 14.2 0.04 0.03

15.3 0.06 0.05

16.6 0.13 0.10

17.8 0.21 0.15

19.7 0.05 0.04

Brilliant II nosZ 50 26.3 0.54 0.38

26.8 0.12 0.09

27.9 0.09 0.07

29.4 0.24 0.17

32.4 0.18 0.13

Quanta 22.2 0.04 0.03

23.7 0.29 0.21

25.7 0.11 0.08

27.6 0.08 0.06

29.6 0.97 0.14

Brilliant II 60 26.2 0.08 0.06

26.8 0.11 0.08

28.2 0.03 0.02

29.6 0.06 0.04

31.6 0.03 0.02

Quanta 25.2 0.04 0.03

26.4 0.24 0.17

27.7 0.11 0.08

29.3 0.06 0.05

31.2 0.44 0.31

Brilliant II qnorB 50 32.7 0.31 0.22

31.5 0.04 0.03

33.1 0.03 0.02

35 0.49 0.35

37.2 0.12 0.09

Quanta 26.2 0.24 0.17

27.7 0.39 0.28

29.7 0.34 0.24

Page 96: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

86

APPENDIX DATA CHAPTER 3 CONTINUED

Appendix material 3.2.6 Calculations for in well concentrations for desired dilution

To calculate “sweet spot”

1. Calculate your 1:2 (or in Jack’s case 1:3) concentrations from the full strength ng µL-1

reading of your samples.

2. Average the 1:3 concentrations (calculated in step 1) to get the Stock I ng µL-1

(e.g. 8.981

ng/uL)

3. Multiply the average Stock I concentration (calculated from step 2) by your dilution

factor, which in our case is 3.

4. Divide the number is step 3 by the desired in well final dilution (e.g. 185.2836).

(8.981*3)/ 185.2836 = 0.142 ng/uL

To convert machine factors to ng/uL

ng/uL = (A*C)/D

A = Stock I average concentration

C = Dilution factor (3)

D = Final dilution factor calculated by PREXCEL-Q

Appendix Table 3.2: Experiment 2: enzyme repeatability performance (n=2) CONTINUED

31.8 0.03 0.02

34.2 0.18 0.13

Brilliant II 55 38.2 1.30 0.92

34.5 0.39 0.28

35.1 0.87 0.62

38.6 1.61 1.14

37.9 1.12 0.79

Quanta 28.4 0.36 0.26

30.3 0.80 0.57

32.1 0.46 0.33

33.5 0.01 0.01

35.9 0.88 0.63

Page 97: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

87

APPENDIX DATA CHAPTER 3 CONTINUED

Appendix Table 3.3: DNA concentrations for Experiment 3: Field and sampling effect August 09 DNA Quantified 2.16.2010 1:2 dilution

Transect Soil Series O.D260 O.D280 260/280 ng/uL ng/uL

K1 Cn 0.964 0.51 1.9 48.2 16.07

Ct 0.951 0.51 1.86 47.5 15.83

Ok 0.622 0.33 1.9 31.1 10.37

W 0.516 0.26 2.01 25.8 8.60

K2 Cn 2.734 1.45 1.88 136.7 45.57

Ct 1.488 0.81 1.84 74.4 24.80

Ok 3.136 1.66 1.89 156.8 52.27

W 0.565 0.30 1.92 28.3 9.43

K3 Cn 0.976 0.53 1.85 48.8 16.27

Ct 0.784 0.41 1.92 39.2 13.07

Ok 0.296 0.15 1.96 14.8 4.93

W 1.174 0.61 1.93 58.7 19.57

C1 Cn 0.63 0.32 1.98 31.5 10.50

Ct 1.388 0.70 1.98 69.4 23.13

Ok 0.932 0.46 2.01 46.6 15.53

W 0.71 0.36 1.95 35.5 11.83

C2 Cn 1.1 0.57 1.94 55 18.33

Ct 1.448 0.76 1.9 72.4 24.13

Ok 1.074 0.57 1.89 53.7 17.90

W 0.966 0.50 1.94 48.3 16.10

C3 Cn 1.278 0.67 1.92 63.9 21.30

Ct 0.44 0.22 1.97 22 7.33

Ok 0.366 0.19 1.9 18.3 6.10

W 1.17 0.60 1.94 58.5 19.50

Average ng/ul of August samples: 17.85

Stock I - October DNA

Quantified 2.5.2010

1:2 dilution

Transect Soil Series 260 280 260/280 260/230 ng/uL ng/uL

K1 Cn 0.839 0.442 1.9 2.21 42 14.00

Ct 0.365 0.196 1.86 2.41 18.2 6.07

Ok 0.235 0.118 2 1.56 11.8 3.93

W 0.12 0.058 2.09 1.27 6 2.00

K2 Cn 0.646 0.329 1.96 1.05 32.3 10.77

Ct 0.325 0.176 1.84 1.38 16.3 5.43

Ok 0.477 0.258 1.85 0.84 23.8 7.93

W 0.216 0.142 1.52 1.12 10.8 3.60

K3 Cn 0.1 0.066 1.53 1.51 5 1.67

Ct 0.461 0.243 1.89 2.12 23.1 7.70

Ok 0.638 0.35 1.82 1.63 31.9 10.63

Page 98: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

88

APPENDIX DATA CHAPTER 3 CONTINUED

Appendix Table 3.3: DNA concentrations for Experiment 3 continued

W 0.2 0.111 1.8 1.63 10 3.33

C1 Cn 0.167 0.126 1.33 0.43 8.4 2.80

Ct 0.216 0.12 1.8 0.75 10.8 3.60

Ok 0.122 0.017 6.96 0.06 6.1 2.03

W 0.503 0.268 1.87 0.78 28.1 9.37

C2 Cn 0.32 0.189 1.7 1 16 5.33

Ct 0.336 0.195 1.72 0.72 16.8 5.60

Ok 0.179 0.088 2.03 0.45 8.9 2.97

W 0.15 0.065 2.33 0.42 7.5 2.50

C3 Cn 0.407 0.224 1.81 1.25 20.3 6.77

Ct 0.267 0.146 1.83 1.75 13.4 4.47

Ok 0.118 0.057 2.07 1.52 5.9 1.97

W 0.247 0.233 1.06 0.76 12.3 4.10

Average ng/ul of October Samples: 5.36

Appendix material 3.2.9 A: Concerns arising from data collection

While in the process of running and analyzing qnorB data from the Kalsow site, I noticed the

melting curves for this gene were very messy (not a single melt curve) and had a very high

melting temperature of 98˚C. Furthermore, efficiencies between runs were very inconsistent (e.g.

41% vs 100%) between plates, but were similar within plates. To troubleshoot, Eppendorf was

contacted and they focused on the cycling conditions and suggested extending the recommended

duration time.

Email correspondence from Dr. Chen Liu, Eppendorf, with Sarah Hargreaves on March 30,

2010:

I would not use 1 second, even if it is recommended. You may want to contact Quanta to see if

there is any special concern. Realplex’s block ramps very fast, I am almost certain that you will

not get 100% denaturation within 1 second. This will greatly affect PCR effificency.

Sarah and I contacted Quanta to discuss changing the denaturation time.

Notes from phone conversation with Quanta Tech Rep, 30 March 2010

Generally, Eppendorf cyclers show “spotty” results with fast cycling

Rep suggests using conservative cycling conditions

o For Fast mixes, this means 5 second denaturation

o For Super mixers, this means 15 or longer denaturation

Page 99: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

89

Rep also suggests using a 3-step program if we are having trouble optimizing results

(again, 3-step programs have been proven to improve reactions with the Eppendorf

cycler)

APPENDIX DATA CHAPTER 3 CONTINUED

Appendix material 3.2.9 A: Concerns arising from data collection continued

When performing a gradient analysis, always perform a 3-step reaction and make the data

collection step the 72C extension step. This is because SYBR fluorescence varies with

pH, and pH changes with temperature resulting in more signal at lower Tm

Buffers and detergent varies between Fast and Super mixes

SYBR is a bit lower in Fast mix, and lower SYBR = nicer melt curves sometimes

Sensitivities in Eppendorf cyclers are known to improve by using white-welled tubes

Eppendorf melt curves are relatively crappy when compared to melt curves using other

cyclers!

Notes from phone conversation with Quanta Tech Rep, 8 April 2010

1. Question: Our efficiency is very variable and low. How can we increase our efficiency

and have tighter reps?

a. Efficiency will be low and variable if you have incomplete denaturation of your

template or your primers are of poor quality.

2. With the standard Super mix is there a difference between a denaturation time of 20

seconds and 30 seconds?

a. With products that are 600bp in length the denaturation step should be 30 seconds

to a minute in time.

b. For a conservative qPCR program with the Quanta SuperMix the initial

denaturation step should be at least five minutes. In this amount of time double

stranded genomic DNA should be fully denatured.

c. There is no difference in the enzymes ability to function if denaturation is 20 or

30 seconds.

Appendix material 3.2.9 B: Dealing with over efficiency

Sarah and I investigated the problem with over efficiency on BitZied Bio, which suggested using

additive to deal with “over” efficient reactions. To further follow up, I contacted Quanta to

discuss using additives to optimize efficiencies.

Notes from phone conversation with Quanta Tech Rep, 9 April 2010

1. Question: Do you have BSA in your SuperMix?

a. There is some BSA in the mix already, but he has used it in his reactions before.

Page 100: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

90

APPENDIX DATA CHAPTER 3 CONTINUED

a. The ability of BSA to help with humic acid and enzyme stability depends on the

company that makes it and its lot number.

b. Dave suggests using a Biotech grade that is non-acetylated

Appendix material 3.2.9 B: Dealing with over efficiency Continued

2. Question: After elongating the denaturation time we now have efficiencies over 110%.

Based on the literature, groups have used T4 gene 32 protein and or BSA. What are your

thoughts about using either of these to help with efficiency?

a. T4 gene will help bring down the efficiency but does not recommend using it.

b. It inhibits PCR after denaturation because it binds to any single stranded DNA

including primers!

3. Our melt peaks are above 85 degrees is this a concern? What does a high melting

temperature mean?

a. High efficiencies may also mean compression between dilution points of your

standard curve

Compression occurs when there is extra carryover of DNA from one dilution to

the next, loss of sample due to the DNA binding to the pipette tip, of due to a

pipetting error.

b. To minimize these effects: check and recheck your pipette settings and dispence

only once (do not mix by pipetting up and down)

c. If melt peaks are reaching 90 degrees your product is not completely denaturing

increase denaturation time

d. If that does not work, try adding either glycerol or DMSO. Both compounds

reduce secondary structures that could inhibit the progress of the polymerase.

Especially useful for GC rich templates.

Page 101: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

91

APPENDIX DATA CHAPTER 4

Appendix Table 4.1 Summary of biological, chemical and physical soil parameters. a values were not transformed for analysis

Soil parameters Value

Abbreviation Description (unit) Mean SE Max Min

WFPSa (vol/vol)

Water-filled Pore Space (%filled by water) 0.84 0.02 1.00 0.13

pHa Soil pH in H2O (vol:vol) 7.4 0.07 8.5 5.9

EOC Extractable Organic carbon ( mg kg-1

dw soil) 64.2 6.6 279.5 0

Tot-C Total Carbon ( % dw soil ) 4.6 0.25 12.9 0.47

NH4+-N Ammonium (mg kg

-1 dw soil) 3.5 0.67 46.12 0.68

NO3--N Nitrate ( mg kg

-1 dw soil ) 3.28 0.85 37.5 0

TIN Total Inorganic Nitrogen ( mg kg-1

dw soil ) 6.8 1.3 70.9 0.85

Avail C: N Available C: N ratio 17.8 1.7 55.0 0

Tot-N Total nitrogen ( % dw soil ) 0.37 0.02 1.02 0.21

C: N Carbon: Nitrogen ratio 12.7 0.20 14.9 1.27

Fe Iron (mg kg-1

dsw) Mehlich-3 extracted 175.7 13.7 730 53

Cu Copper (mg/kg-1

dsw) Mehlich-3 extracted 8.40 0.34 16.0 3.5

16S rDNA rRNA Eubacteria gene copy number (g-1

dw soil) 2.56x1010

3.41x109 2.00x10

11 7.79x10

7

qnorB qnorB gene copy number (g-1

dw soil) 7.1x104 1.8x10

4 9.45x10

5 2.70x10

3

nosZ nosZ gene copy number (g-1

dw soil) 1.27x107 2.1x10

6 1.02x10

8 4.97x10

5

qnorB:16S rDNA qnorB:16S rDNA gene copy number ratio (%) 0.001 0.0007 0.05 3.0 x10-5

nosZ: 16S rDNA nosZ:16S rDNA gene copy number ratio (%) 0.06 0.09 0.62 0.01

qnorB:nosZ qnorB:nosZ gene copy number ratio (%) 0.97 0.29 18.7 0.03

Page 102: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

92

APPENDIX DATA CHAPTER 4 CONTINUED

Appendix Figure 4.1 . Schematic diagram of soil catena at Kalsow Prairie Preserve and adjacent

agricultural ecosystems. Changes in topography and corresponding soil series provide a natural

environmental gradient for examining the biogeography of denitrifying bacterial communities.

Page 103: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

93

APPENDIX DATA CHAPTER 4 CONTINUED

Means and standard errors are given. a not transformed

N=18 Comparison within soil type between land-uses. Different capital letters following the

mean values indicate a significant difference. Italicized letters indicate a marginal difference

(ANOVA P< 0.05, Tukey test)

Supplementary Table 4.2a: Selected properties of surface soil (0-10 cm) within soil type

Soil property Clarion Webster Canisteo Okoboji

WFPSa 0.81

(0.01) 0.84

(0.02) 0.85

(0.02) 0.86 (0.01)

Soil pHa

6.92 B

(0.11) 7.50 C

(0.08) 8.10 A

(0.07) 7.10 B

(0.14)

EOC 41.4 A

(7.4) 51.4 A

(9.0) 110.3 B

(14.0) 50.3 A

(14.5)

TC 4.42 B

(0.30) 5.35 A B

(0.52) 5.74 A

(0.43) 5.60 A B

(0.67)

NH4+ 6.20 (2.6) 2.80

(0.35) 2.64 (0.39) 2.40 (0.37)

NO3- 2.60

B (1.7) 1.30

B (˂0.01) 3.40

A (1.0) 5.90

B (2.8)

TIN 8.80 (4.1) 4.10 (0.58) 6.01 (1.1) 8.30 (2.8)

Avail C:N 13.4 B

(2.6) 15.5 B

(2.5) 26.1 A

(2.5) 16.4 B (3.4)

TN 0.36 (0.02) 0.42 (0.04) 0.43 (0.03) 0.46 (0.05)

C:N 12.2 B

(0.23) 12.8 C

(0.21) 13.3 A

(0.21) 12.0 B C

(0.21)

Iron 199.2 BC

(36.0) 155.4 C

(19.0) 93.8 AC

(7.9) 254.2 B

(24.7)

Copper 6.90 C

(0.63) 8.25 A

(0.69) 8.72 A

(0.63) 9.72 B

(0.68)

Page 104: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

94

APPENDIX DATA CHAPTER 4 CONTINUED

Means and standard errors are given. a not transformed

N=24 Comparison within soil type between land-uses. Different capital letters following the

mean values indicate significant difference. Italicized letters indicate a marginal difference

(ANOVA P< 0.05, Tukey test)

Supplementary Table 4.2b: Selected properties of surface soil (0-10 cm) between land management

Under two land managements

Soil property Agriculture Prairie

WFPSa (vol/vol) 0.77

A (0.02) 0.92

B (0.01)

Soil pHa

7.45 (0.10) 7.34 (0.11)

EOC 36.7 A

(5.0) 90.8 B

(10.5)

TC 3.87 A

(0.93) 6.67 B

(0.36)

NH4+ 3.45

A (0.13) 3.51

B (0.32)

NO3- 4.96

A (1.47) 1.61

B (0.80)

TIN 8.4 A

(2.37) 5.2 B

(0.88)

Avail C:N 12.9 A

(2.2) 22.7 B

(2.4)

TN 0.31 A

(0.01) 0.53 B

(0.03)

Soil C:N 12.6 (0.19) 12.6 (0.35)

Iron (mg/g-1

dsw) 174.4 (19.3) 176.9 (19.8)

Copper (mg/g1dsw) 9.40

(0.51) 7.40

(0.002)

Page 105: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

95

APPENDIX DATA CHAPTER 4 CONTINUED

Means and standard errors are given. a not transformed

N=36 Comparison within soil type between land-uses. Different capital letters following the

mean values indicate significant difference. Italicized letters indicate a marginal difference

(ANOVA P< 0.05, Tukey test)

Appendix Table 4.2c: Selected properties of surface soil (0-10 cm) by sampling date

Soil property May August October

WFPSa

0.87 A

(0.03) 0.74 B (0.04) 0.92

A (0.03)

Soil pHa

7.15 A

(0.11) 7.40 B

(0.14) 7.70

C (0.11)

EOC 58.9

A (10.8) 79.8

B (14.8) 54.3

A (6.9)

TC 5.15

(0.32) 5.70

(0.58) 4.96 (0.37)

NH4+

6.00 A

(1.91) 2.84 B (0.36) 1.71

C (0.16)

NO3-

6.40 A

(2.10) 2.61 A

(1.14) 0.87 C (0.62)

TIN 12.3

A (3.3) 5.5

B (1.24) 2.6

C (0.66)

Avail C: N 10.9

A (2.53) 18.3

B (2.50) 24.3

B (3.20)

TN 0.39

A (0.02) 0.47

B (0.04) 0.39

A (0.02)

Soil C:N 13.2

A (0.13) 11.9

B (0.22) 12.6

A (0.51)

Iron 140.1

A (15.9) 204.0

B (32.0) 182.8

B (19.2)

Copper 7.7

A (0.59) 10.5

B (0.49) 6.9

B (0.44)

Page 106: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

96

APPENDIX DATA CHAPTER 4 CONTINUED

Means and standard errors are given. a not transformed

N= 12 Comparison between land management by sampling date. Different capital letters

following the mean values indicate significant difference between land managements. Lower

case letters indicate significant differences within land managements by sample date. Italicized

letters indicate a marginal difference (ANOVA P< 0.05, Tukey test)

Supplementary Table 4.3a Soil properties of surface soil (0-10cm) by sample date by land management

May August October

Soil Property Agriculture Prairie Agriculture Prairie Agriculture Prairie

WFPSa

0.78 (0.02)

0.96

(0.03)

0.65

(0.02) 0.84

(0.03) 0.87

(0.02) 0.96 (0.05)

pHa

7.03 a

(0.167)

7.26 a

(0.15)

7.64 b

(0.14) 7.11 a (0.21) 7.69

b (0.14) 7.64

b (0.17)

EOC (mg/kg) 30.6 A

(7.7) 86.6 B ab

(16.9) 40.2 A

(9.5) 119.5 B a

(23.3) 39.5 ( 9.1) 66.4 b

(9.4)

TC 3.92 A

(0.18) 6.38 B

(0.36)

3.69 A

(0.15) 7.70 B

(0.79) 4.00 A

(0.16) 5.92 B

(0.62)

NH4+(mg/kg) 7.44

(3.8)

4.52

(0.66)

1.80

(0.28)

3.96 (0.52) 1.21

(0.15) 2.20

(0.15)

NO3-(mg/kg)

11.9

A a (3.6)

0.84

B ab (0.40)

1.34

ab (0.29)

3.90

a (2.3) 1.65

A b (1.2) 0.10

B b (0.09)

TIN 19.3 A (5.97) 5.36

B ac (0.78) 3.14

A (0.34) 7.90

B a (2.30) 2.86

A (1.34) 2.30

B b (0.18)

Avail C: N 3.9 (1.30)

17.9 (4.02) 14.3 (2.80) 22.3 (3.96) 20.6 (4.70) 27.9 (4.30)

TN 0.30 A

(0.02) 0.48 B a

(0.03) 0.32 A

(0.012) 0.63 B b

(0.06) 0.31 A

(0.01) 0.47 B a

(0.03)

Soil C:N 12.9 (0.21)

13.4 (0.14) 11.6

(0.412) 12.1

(0.17) 13.0

(0.17) 12.2 (1.00)

Fe3+

129.6 (19.0) 150.7

(26.2) 201.1

(51.9) 207.0

(40.0) 192.5

(14.7) 173.0

(36.1)

Cu 8.1 (0.68)

7.3 (1.0) 12.3 (0.61)

8.8 (0.35)

7.9

(0.74) 6.0

(0.31)

Page 107: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

97

97

Means and standard errors are given.

N=6 Comparison between soil type over the season.

Different capital letters following the mean values indicate significant difference between soil types within a season. Different lower

case letters indicate significant differences within soil type by sample date. Italicized letters indicate a marginal difference (ANOVA

P< 0.05, Tukey test).

Appendix Table 4.3b Soil properties of surface soil (0-10cm) by sampling date by soil type

N=6 May August October

Soil

Properties Clarion Webster Canisteo Okoboji Clarion Webster Canisteo Okoboji Clarion Webster Canisteo Okoboji

WFPS a

0.81

(0.04)

0.88

(0.06)

0.91

(0.06)

0.88

(0.06)

0.72

(0.04)

0.64 a

(0.06)

0.73

(0.05)

0.86

(0.13)

091

(0.04)

0.98 b

(0.09)

0.92

(0.02)

0.85

(0.03)

pH a

6.50 A a

(0.12)

7.39 B

(0.10)

7.80 B

(0.067)

6.88 A a

(0.14)

6.98 A

(0.15)

7.37 A

(0.13)

8.23 B

(0.09)

6.95 A

(0.29)

7.28 A b

(0.147)

7.67 A

(0.184)

8.28 B

(0.062)

7.44 A b

(0.22)

EOC 34.5

(6.14)

61.0

(20.1)

104.1

(16.1)

36.0

(27.4)

53.2

(19.2)

53.5 B

(15.7)

145.4

(36.3)

67.1

(33.7)

40.0

(17.7)

40.1

(11.1)

81.4

(6.4)

47.8

(11.1)

TC 4.37

(0.54)

4.87

(0.55)

5.63

(0.72)

5.73 a

(0.75)

4.29

(0.58)

6.04

(1.42)

5.53

(0.78)

6.93 a

(1.56)

4.59

(0.53)

5.13

(0.55)

6.06

(0.87)

4.06 b

(0.86)

NH4+

14.5 a

(6.8)

3.7

(0.67)

3.3

(0.49)

2.4

(0.41)

2.3

(0.37)

2.8

(0.55)

3.0

(0.97)

3.1

(0.96)

1.7 b

(0.33)

1.9

(0.41)

1.7

(0.29)

1.5

(0.27)

NO3- 7.0

(4.3)

1.7

(0.87)

5.2

(1.7)

11.5

(6.9)

0.67

(0.25)

2.0

(0.92)

1.7

(0.44)

6.2

(4.5)

0.05

(0.03)

0.11

(0.08)

3.2

(2.4)

0.12

(0.07)

TIN 21.6

(11.0)

5.40

(0.59)

8.51

(1.42)

13.9

(6.71)

3.07

(0.31)

4.86

(1.30)

4.58

(1.03)

9.29 a

(4.63)

1.74

(0.31)

1.98

(0.39)

4.93

(2.51)

1.67 b

(0.25)

Avail C:N 5.30

(1.64)

14.3

(5.92)

15.5

(4.52)

8.50

(6.75)

17.1

(3.82)

11.7

(3.22)

31.5

(5.10)

12.8

(3.22)

17.9

(7.92)

20.4

(4.36)

31.0

(6.97)

27.9

(5.97)

TN 0.34

(0.04)

0.38

(0.04)

0.41

(0.05)

0.43

(0.05)

0.38

(0.04)

0.49

(0.11)

0.45

(0.07)

0.57 a

(0.12)

0.34

(0.10)

0.39

(0.04)

0.44

(0.06)

0.37 b

(0.09)

Soil C:N 12.8

(0.15)

12.9

(0.16)

13.8

(0.25)

13.2

(0.30)

11.2

(0.37)

12.3

(0.57)

12.3

(0.31)

11.9

(0.40)

12.7

(0.19)

13.2

(0.14)

13.7

(0.18)

10.9

(1.9)

Fe3+

140.3

AB

(2.6)

98.7 A

(10.0)

74.2 A

(2.6)

247.3 B

(28.3)

245.4 B

(98.0)

175.8 AB

(94.2)

87.5 A

(11.4)

307.5 B

(36.9)

211.9

(46.1)

191.7

(12.7)

119.6

(17.0)

208.0

(55.4)

Cu 5.7

AC a

(0.94)

6.6 C a

(0.44)

7.6 B C

(0.76)

10.8 B

(1.3)

9.6 b

(0.82)

11.4 b

(1.1)

10.8

(1.0)

10.3

(1.1)

5.4 A a

( 0.54)

6.8 AB a

(0.62)

7.6 AB

(1.0)

8.0 B

(1.0)

AP

PE

ND

IX D

AT

A C

HA

PT

ER

4 C

ON

TIN

UE

D

Page 108: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

98

98

Means and standard errors are given. a not transformed

N= 9 Comparison between land management and soil type. Different capital letters following the mean values indicate

significant difference within soil between land managements. Lower case letters indicate differences between soils within land

managements . Italicized letters indicate a marginal difference (ANOVA P< 0.05, Tukey test

AppendixTable 4.4 Soil properties of surface soil (0-10cm) by soil type by land management

N=9 Clarion Webster Canisteo Okoboji

Soil Properties Agriculture Prairie Agriculture Prairie Agriculture Prairie Agriculture Prairie

WFPSa 0.80 (0.04) 0.83 (0.04) 0.75 (0.05) 0.95 (0.05) 0.78 (0.03) 0.93 (0.04) 0.74

A (0.05) 0.99

B (0.09)

pHa 6.9

b (0.18) 6.9

b (0.14) 7.5

c (0.11) 7.5

c (0.13) 8.1

a (0.12) 8.1

a (0.07) 7.4

b c (0.16) 6.8

b (0.18)

EOC (mg/kg) 20.0 (3.50) 64.5 (10.0) 27.5 (7.60) 77.4 (11.6) 74.3

(7.90) 148.2 (20.8) 28.7

(8.70) 73.2 (26.4)

TC 3.40 (0.26) 5.44 (0.23) 3.90 (0.13) 6.80 (0.78) 4.02

(0.09) 7.50 (0.21) 4.20 (0.13) 6.96

(1.20)

NH4+(mg/kg) 8.23

(0.23) 4.17

(1.90) 2.30 (0.59) 3.31

(0.34) 1.80 (0.24) 3.51 (0.63) 1.52

(0.27) 3.27 (0.58)

NO3-(mg/kg)

5.03 (3.00) 0.17 (0.10) 1.62 (0.63) 0.91 (0.64) 5.30 (1.70) 1.45 (0.51) 7.90 (4.80) 3.90 (3.10)

TIN 13.3 (8.1) 4.33 (0.99) 3.92 (0.76) 4.22 (0.92) 7.10 (1.90) 4.90 (0.95) 9.41(4.86) 7.16 (3.17)

Avail C:N 6.44 (1.50) 20.43 (5.22) 9.89 (3.30) 21.03 (3.39) 17.9 (4.72) 34.1 (3.82) 17.4 (5.79) 15.4 (4.81)

TN 0.29 (0.02) 0.43 (0.02) 0.31 (0.01) 0.53

(0.06) 0.30 (0.007) 0.57

(0.02) 0.34 (0.01) 0.57

(0.07)

Soil C:N 11.9 a (0.39) 12.6 (0.22) 12.8

ab (0.35) 12.8

(0.27) 13.4

b (0.29) 13.1

(0.32) 12.2

a (0.30) 11.7

(1.3)

Fe3+

(mg/kg) 238.8 (63.1) 159.6

(33.4) 136.9

(18.9) 173.8

(32.9) 96.9

(15.3) 90.6 (5.4) 224.9 (12.5) 238.6

(47.2)

Cu (mg/kg) 8.00 (0.95) 5.80 (0.89) 9.11

(3.50) 7.40

(0.72) 9.70

(1.10) 7.80

(0.56) 10.8

(0.80) 8.61 (1.00)

AP

PE

ND

IX D

AT

A C

HA

PT

ER

4 C

ON

TIN

UE

D

Page 109: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

99

99

Appendix Table 4.6 : Pearson Correlation ( r ) between qnorB, nosZ, qnorB:16S, nosZ:16S , and qnorB:nosZ abundances

and soil parameters in full model

Parameters N= 72

Abbreviation : Description: qnorB nosZ qnorB:16S nosZ:16S qnorB:nosZ

qnorB (copy g-1

dsw) Nitric oxide reducatase - 0.54*** - - -

nosZ (copy g-1

dsw) Nitrous oxide reductase 0.54*** - - - -

WFPS Water-filled pore space 0.02 -0.13 -0.01 -0.10 0.12

pH Soil pH (vol soil: vol H2O) -0.10 0.04 -0.07 0.10 -0.11

EOC (mg/kg) Extractable Organic Carbon 0.06 0.34** -0.08 0.31** -0.26*

TC Total Carbon 0.10 0.17 -0.05 0.05 -0.06

NH4+(mg/kg) Ammonium 0.10 0.20 -0.05 0.05 -0.12

NO3-(mg/kg)

Nitrate 0.04 0.17 0.18 0.29* -0.06

TIN Total Inorganic Nitrogen 0.14 0.24 -0.08 0.05 -0.10

Avail C:N Available C:N -0.14 0.07 -0.21 0.13 -0.22

TN Total Nitrogen 0.05 0.16 -0.08 0.06 -0.10

C:N Carbon: Nitrogen -0.13 -0.35** 0.13 -0.17 0.17

Fe3+

(mg/kg) Iron (Mehlich-3) 0.10 -0.14 0.10 -0.16 0.19

Cu (mg/kg) Copper (Mehlich-3) 0.19 0.25* 0.13 0.23 -0.07

Darker shaded gray: statistically significant P<0.05*, P<0.01**, P<0.001***

Values that are under lined and shaded light gray: marginally significant P≤ 0.09

Appendix Table 4.5 Primers for molecular assays

gene Forward

Primer

Forward Primer Sequence Reverse

Primer

Reverse Primer Sequence Amplicon

Length bp

Annealing

temperature

Source

nosZ nosZ1F WCSYTGTTCMTCGACAGCCAG nosZ1R ATGTCGATCARCTGVKCRTTYTC 259 60˚C Henry et al.. 2006

qnorB qnorB2F GGNCAYCARGGNTAYGA qnorB5R ACCCANAGRTGNACNACCCACCA 262 56˚C Braker & Tiedje

2003

16s Eub338 ACTCCTACGGGAGGCAGCAG Eub518 ATTACCGCGGCTGCTGG 200 60˚C Fierer et al.. 2005

AP

PE

ND

IX D

AT

A C

HA

PT

ER

4 C

ON

TIN

UE

D

Page 110: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

100

100

Appendix Table 4.7 Pearson Correlation ( r ) between qnorB, nosZ, qnorB:16S, nosZ:16S , and qnorB:nosZ abundances and soil parameters by land use

N= 36 Agriculture Prairie

Parameters qnorB nosZ qnorB:16S nosZ:16S qnorB:nosZ qnorB nosZ qnorB:16S nosZ:16S qnorB:nosZ

qnorB - 0.55** - - - - 0.52** - - -

nosZ 0.55** - - - - 0.52** - - - -

WFPS -0.20 -0.44** -0.09 -0.23 0.24 0.13 -0.04 0.25 0.04 0.18

pH 0.12 0.23 0.05 0.23 -0.11 -0.30 -0.11 -0.21 -0.04 -0.16

EOC -0.07 0.18 -0.08 0.36* -0.26 0.10 0.41* 0.04 0.46** -0.20

TC 0.02 -0.23 0.12 -0.05 0.25 0.10 0.25 0.12 0.34* 0.02

NH4+ -0.03 0.15 -0.01 -0.02 -0.06 0.28 0.40* 0.12 0.35* -0.07

NO3-

0.04 -0.15 0.05 0.10 -0.19 0.18 0.37* 0.19 0.47** -0.08

TIN 0.13 0.005 0.02 0.02 0.14 0.14 0.37* 0.11 0.32 -0.07

Avail C:N -0.12 0.02 -0.14 0.23 -0.16 -0.26 -0.01 -0.18 0.10 -0.20

TN 0.005 0.13 0.01 0.03 0.14 -0.03 0.21 0.15 0.37* -0.07

C:N -0.15 -0.35* 0.20 -0.11 0.19 -0.14 -0.44** 0.10 -0.25 0.21

Fe3+

-0.19 -0.24 -0.15 -0.18 0.04 0.37* -0.03 0.35* -0.15 0.36*

Cu 0.27 0.46** 0.11 0.22 -0.17 0.18 0.18 0.04 0.23 -0.09

AP

PE

ND

IX D

AT

A C

HA

PT

ER

4 C

ON

TIN

UE

D

Page 111: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

101

101

Appendix Table 4.8 Pearson Correlation ( r ) between qnorB, nosZ, qnorB:16S, nosZ:16S, and qnor:nosZ abundances

and soil parameters by soil type

N= 18 Clarion Webster

Parameters: qnorB nosZ qnorB:16S nosZ:16S

qnorB:nos

Z qnorB nosZ qnorB:16S nosZ:16S qnorB:nosZ

qnorB - 0.45 - - - - 0.72*** - - -

nosZ 0.45 - - - - 0.72*** - - -

WFPS 0.19 -0.23 0.02 -0.39 0.37 -0.17 -0.34 0.20 -0.15 0.26

pH -0.05 -0.05 -0.28 -0.46 -0.02 -0.23 -0.03 -0.27 0.03 -0.25

EOC 0.05 0.36 0.02 -0.28 -0.19 -0.26 0.09 -0.28 0.34 -0.44

TC 0.26 0.21 -0.09 -0.37 0.14 -0.09 0.12 -0.17 0.22 -0.27

NH4+ -0.12 -0.01 0.12 -0.01 -0.13 -0.10 0.07 -0.18 0.10 -0.21

NO3-

-0.29 -0.18 0.02 0.41 -0.20 0.53* 0.57* 0.03 0.25 -0.12

TIN 0.31 0.37 -0.02 -0.27 0.24 -0.10 0.19 -0.22 0.35 -0.39

Avail C:N 0.14 0.32 -0.06 -0.32 0.17 -0.57* -0.29 -0.32 0.05 -0.31

TN 0.32 0.37 -0.01 -0.27 0.10 -0.26 0.08 -0.23 0.38 0.42

C:N 0.20 -0.03 0.006 -0.42 0.17 -0.01 -0.34 0.23 -0.44 0.45

Fe3+

0.14 0.12 0.07 0.09 0.07 -0.13 -0.19 -0.06 -0.26 0.10

Cu 0.31 0.44 0.33 0.44 0.04 0.42 0.38 0.15 0.20 0.01

Canisteo Okoboji

qnorB nosZ qnorB:16S nosZ:16S qnorB:nosZ qnorB nosZ qnorB:16S nosZ:16S qnorB:nosZ

qnorB - 0.37 - - - - 0.59** - - -

nosZ 0.37 - - - - 0.59** - - - -

WFPS -0.34 -0.48* -0.09 -0.32 0.23 0.17 0.30 -0.26 0.11 -0.28

pH 0.25 0.36 -0.16 0.14 -0.18 -0.28 -0.24 -0.25 -0.21 -0.01

EOC -0.05 0.46 -0.60** 0.14 -0.50* 0.23 0.33 0.13 0.30 0.03

TC -0.18 -0.04 -0.38 -0.35 -0.09 0.25 0.45 -0.001 0.22 -0.15

NH4+ -0.09 0.14 -0.37 -0.09 -0.21 0.56* 0.52* 0.27 0.28 0.08

NO3-

-0.02 0.02 0.14 0.25 0.03 0.04 0.08 0.35 0.13 0.22

TIN -0.09 0.07 -042 -0.31 -0.14 0.29 0.53* 0.07 0.30 -0.14

Avail C:N -0.05 0.31 -0.50* -0.02 -0.35 -0.23 -0.11 -0.12 -0.07 -0.16

TN 0.08 -0.10 -0.43 -0.30 0.15 -0.09 0.29 0.12 0.28 0.03

C:N -0.39 -0.59** 0.33 -0.14 0.31 -0.37 -0.12 -0.23 -0.33 -0.02

Fe3+

0.01 -0.42 0.42 -0.25 0.43 0.33 0.15 0.28 0.22 -0.15

Cu 0.39 0.46 -0.12 0.17 -0.18 -0.10 0.18 -0.20 0.30 -0.37

AP

PE

ND

IX D

AT

A C

HA

PT

ER

4 C

ON

TIN

UE

D

Page 112: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

102

102

Appendix Table 4.9a Pearson Correlation ( r ) between qnorB, nosZ, qnorB:16S, qnosZ:16S , and qnor:nosZ abundances and

soil parameter under agricultural land management by soil type

Agriculture

N= 9 Clarion Webster

Parameters qnorB nosZ qnorB:16S nosZ:16S qnorB:nosZ qnorB nosZ qnorB:16S nosZ:16S qnorB:nosZ

qnorB - 0.19 - - - - 0.77* - - -

nosZ 0.19 - - - - 0.77* - - - -

WFPS 0.43 -0.33 -0.02 -0.16 0.56 -0.49 -0.76* 0.08 -0.54 0.46

pH 0.11 0.31 -0.29 -0.29 0.11 0.34 -0.33 0.04

EOC -0.27 0.47 -0.24 0.29 -0.46 -0.27 0.20 -0.58 0.13 -0.69*

TC 0.38 0.07 -0.28 0.12 0.36 0.27 0.17 0.31 0.26 0.15

NH4+ -0.22 -0.24 0.12 -0.04 -0.11 -0.23 -0.13 -0.19 -0.10 -0.12

NO3-

-0.54 -0.01 -0.03 0.15 -0.54 0.53 0.52 0.09 0.17 -0.03

TIN 0.39 0.44 -0.27 0.17 0.20 0.33 0.49 0.23 0.73* -0.28

Avail C:N 0.17 0.60 -0.18 0.29 -0.08 -0.45 -0.17 -0.46 -0.13 -0.38

TN 0.40 0.44 -0.28 0.12 0.20 0.33 0.50 0.24 0.73* -0.28

C:N -0.06 -0.84** 0.13 -0.11 0.29 0.03 -0.27 0.11 -0.73* 0.45

Fe3+

0.11 0.42 -0.03 0.11 -0.06 -0.26 -0.31 -0.05 -0.21 0.10

Cu 0.52 0.15 0.38 0.07 -0.07 0.71* 0.70* 0.31 0.51 -0.05

Agriculture

Canisteo Okoboji

qnorB nosZ qnorB:16S nosZ:16S qnorB:nosZ qnorB nosZ qnorB:16S nosZ:16S qnorB:nosZ

qnorB - 0.61 - - - - 0.67* - - -

nosZ 0.61 - - - - 0.67* - - - -

WFPS -0.28 -0.64 0.53 -0.41 0.58 -0.85** -0.46 -0.84** -0.14 -0.38

pH 0.26 0.52 -0.45 0.25 -0.05 0.02 -0.26 0.15

EOC -0.10 0.44 -0.88** 0.07 -0.64 -0.26 -0.38 -0.23 0.07 0.19

TC -0.26 -0.74* 0.65 -0.52 0.71* -0.62 -0.52 -0.37 -0.31 -0.03

NH4+ -0.26 0.03 -0.43 -0.08 -0.25 0.48 0.18 0.47 0.11 0.32

NO3-

-0.22 -0.33 0.06 -0.35 0.22 0.04 -0.04 0.04 -0.35 0.10

TIN 0.54 -0.20 0.60 -0.55 0.70* -0.09 0.09 -0.20 0.13 -0.21

Avail C:N 0.22 0.42 -0.31 0.25 -0.34 -0.33 -0.24 -0.28 0.24 -0.06

TN 0.54 -0.19 0.59 -0.55 0.70* -0.10 0.09 -0.20 0.13 -0.21

C:N -0.84** -0.51 -0.01 0.11 -0.07 -0.67 -0.77 -0.23 -0.67 0.23

Fe3+

-0.14 -0.60 0.62 -0.40 -0.64 -0.15 0.07 -0.11 0.61 -0.28

Cu 0.75* 0.62 -0.37 -0.20 -0.15 0.30 0.81** -0.28 0.83** -0.69

AP

PE

ND

IX D

AT

A C

HA

PT

ER

4 C

ON

TIN

UE

D

Page 113: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

103

103

Appendix Table 4.9b Pearson Correlation ( r ) between qnorB, nosZ, qnorB:16S, qnosZ:16S , and qnor:nosZ abundances and soil parameter under

prairie land management by soil type

Prairie

N= 9 Clarion Webster

Parameters qnorB nosZ qnorB:16S nosZ:16S qnorB:nosZ qnorB nosZ qnorB:16S nosZ:16S qnorB:nosZ

qnorB - 0.79** - - - - 0.73* - - -

nosZ 0.79** - - - - 0.73* - - - -

WFPS -0.21 -0.20 -0.25 -0.41 -0.06 0.28 -0.15 0.62 -0.07 0.41

pH -0.35 -0.40 -0.30 -0.64 -0.50 -0.28 -0.23 0.01

EOC 0.38 0.24 0.28 -0.06 0.27 0.10 0.08 0.36 0.50 -0.08

TC 0.41 0.12 0.51 -0.18 0.51 0.34 0.27 0.11 0.14 -0.03

NH4+ 0.12 0.16 0.15 0.41 -0.03 0.76* 0.56 0.36 0.37 -0.32

NO3-

-0.03 -0.12 0.20 0.18 0.11 0.45 0.64 -0.20 0.44 0.44

TIN 0.54 0.30 0.54 0.05 0.45 0.34 0.37 0.02 0.28 -0.34

Avail C:N 0.13 0.06 0.02 -0.39 0.13 -0.59 -0.55 0.19 0.28 -0.33

TN 0.54 0.29 0.57 0.03 0.45 0.09 0.25 0.03 0.51 -0.34

C:N -0.48 -0.50 -0.32 -0.56 -0.07 -0.04 -0.43 0.44 -0.45 -0.54

Fe3+

0.25 0.06 0.30 -0.19 0.33 0.12 -0.10 0.09 -0.42 0.35

Cu 0.70* 0.68* 0.41 0.47 0.18 -0.05 -0.08 0.36 0.36 -0.13

Prairie

Canisteo Okoboji

qnorB nosZ qnorB:16S nosZ:16S qnorB:nosZ qnorB nosZ qnorB:16S nosZ:16S qnorB:nosZ

qnorB - 0.13 - - - - 0.39 - - -

nosZ 0.13 - - - - 0.39 - - - -

WFPS -0.33 -0.63 0.45 -0.08 0.31 0.31 0.79* -0.16 0.49 -0.41

pH 0.26 0.04 0.24 0.02 -0.09 -0.17 -0.18 -0.66

EOC 0.21 0.67* -0.21 0.57 -0.42 0.06 0.76* 0.16 0.53 -0.08

TC 0.17 -0.53 -0.41 -0.59 0.55 -0.21 0.67 -0.21 0.67 -0.03

NH4+ 0.36 0.23 0.24 0.20 0.04 0.32 0.75* 0.01 0.72* -0.32

NO3-

-0.09 0.36 -0.26 0.39 -0.36 0.41 0.62 0.68 0.70* 0.44

TIN 0.17 0.11 -0.09 -0.14 0.02 -0.04 0.70* -0.01 0.65 -0.34

Avail C:N -0.19 0.11 -0.33 0.03 -0.21 -0.22 0.09 -0.23 0.17 -0.33

TN 0.17 0.10 -0.08 -0.15 0.02 -0.36 -0.18 0.07 0.64 -0.34

C:N -0.04 -0.68* 0.56 -0.41 0.55 -0.28 0.67 -0.46 0.04 -0.54

Fe3+

0.38 0.03 0.32 -0.06 0.22 0.48 0.21 0.39 0.19 0.34

Cu -0.29 0.34 -0.35 0.50 -0.47 0.03 0.01 -0.09 0.05 -0.13

AP

PE

ND

IX D

AT

A C

HA

PT

ER

4 C

ON

TIN

UE

D

Page 114: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

104

104

Appendix Table 4.10a Pearson Correlation ( r ) between qnorB, nosZ abundances and soil parameter by month full model

N= 9 May August October

Parameters qnorB nosZ qnorB nosZ qnorB nosZ

qnorB - 0.29 - 0.31 - 0.58**

nosZ 0.29 - 0.31 - 0.58** -

WFPS 0.31 0.51* 0.11 0.25 0.29 0.07

pH 0.04 0.17 -0.19 0.24 -0.10 -0.06

EOC 0.03 0.45* 0.07 0.46* -0.16 -0.08

TC 0.05 0.47* 0.04 0.09 0.14 0.18

NH4+ -0.06 0.29 0.15 0.17 0.11 0.07

NO3-

-0.28 -0.28 0.33 0.24 -0.1 -0.05

TIN 0.01 0.43* 0.05 0.07 0.15 0.13

Avail C:N 0.05 0.39 -0.29 0.29 -0.33 -0.25

TN 0.005 0.44* -0.14 -0.17 0.13 -0.16

C:N 0.26 0.31 0.06 -0.11 0.08 0.16

Fe3+

0.25 -0.15 -0.01 -0.30 -0.03 -0.27

Cu -0.01 -0.27 0.04 0.05 -0.37 -0.37

Appendix Table 4.10b Pearson Correlation ( r ) between qnorB:16S, qnosZ:16S , and qnor:nosZ abundances and soil parameter by month full model

N= 9 May August October

Parameters qnorB:16S nosZ:16S qnorB:nosZ qnorB:16S nosZ:16S qnorB:nosZ qnorB:16S nosZ:16S qnorB:nosZ

WFPS -0.10 0.13 -0.16 -0.08 0.17 -0.12 0.23 -0.04 0.29

pH 0.27 0.15 0.09 -0.30 0.33 -0.37 0.04 0.21 -0.17

EOC 0.04 0.30 -0.17 -0.25 0.39 -0.35 0.07 0.03 -0.13

TC -0.40 0.04 -0.31 0.02 0.05 0.04 0.04 -0.09 0.01

NH4+ -0.37 0.001 -0.26 -0.01 0.05 -0.03 0.07 -0.34 0.08

NO3-

0.23 -0.002 0.17 0.16 0.22 0.06 0.19 0.46* -0.08

TIN -0.45* 0.01 -0.33 0.03 0.05 0.01 0.03 -0.15 0.06

Avail C:N -0.02 0.35 -0.24 -0.45* 0.34 -0.50* -0.15 -0.09 -0.19

TN -0.45* 0.01 -0.33 0.07 0.17 0.001 0.005 -0.13 0.02

C:N 0.14 0.21 -0.04 0.13 -0.10 0.14 0.24 0.23 -0.05

Fe3+

-0.16 -0.14 -0.005 0.23 -0.18 0.26 0.12 0.23 0.27

Cu 0.11 0.02 0.06 -0.06 -0.10 -0.01 0.43* 0.08 0.20

AP

PE

ND

IX D

AT

A C

HA

PT

ER

4 C

ON

TIN

UE

D

Page 115: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

105

105

Appendix Table 4.11a Pearson Correlation ( r ) between qnorB and nosZ abundances and soil parameterby month under agricultural management

N= 9 May August October

Parameters qnorB nosZ qnorB nosZ qnorB nosZ

qnorB - 0.13 - 0.20 - 0.58*

nosZ 0.13 - 0.20 - 0.58* -

WFPS 0.25 0.28 -0.65* 0.02 0.58* 0.13

pH 0.31 0.10 0.25 0.48 -0.17 0.09

EOC 0.01 0.45 0.02 0.65* -0.22 -0.31

TC -0.29 -0.03 0.12 0.01 0.35 -0.24

NH4+ -0.35 0.03 0.19 0.30 0.15 -0.27

NO3- -0.16 -0.10 0.25 0.54 -0.06 0.24

TIN -0.28 -0.07 0.10 0.13 0.33 -0.37

Avail C:N -0.13 0.46 -0.24 0.22 -0.40 -0.49

TN -0.29 -0.07 0.09 0.11 0.33 -0.36

C:N 0.09 0.15 0.08 -0.13 0.13 0.32

Fe3+ -0.34 -0.40 -0.51 -0.40 -0.14 -0.50

Cu 0.08 -0.14 0.30 0.45 -0.09 -0.24

Appendix Table 4.11b Pearson Correlation ( r ) between qnorB:16S, qnosZ:16S , and qnor:nosZ abundances and soil parameter by month under

agricultural management

N= 9 May August October

Parameters qnorB:16S nosZ:16S qnorB:nosZ qnorB:16S nosZ:16S qnorB:nosZ qnorB:16S nosZ:16S qnorB:nosZ

WFPS 0.08 0.05 0.02 -0.61* -0.04 -0.48 0.29 -0.12 -0.30

pH 0.33 0.03 0.19 0.02 0.63* -0.23 0.12 0.35 -0.27

EOC -0.08 0.35 -0.30 -0.31 0.75** -0.55 0.15 0.06 -0.07

TC -0.39 -0.05 -0.22 0.15 0.09 0.07 0.43 -0.03 0.59*

NH4+ -0.46 0.02 -0.31 -0.10 0.13 -0.13 0.25 -0.44 0.37

NO3- -0.18 -0.07 -0.07 -0.29 0.11 -0.28 0.32 0.46 -0.24

TIN -0.41 -0.10 -0.19 0.07 0.22 -0.03 0.42 -0.15 0.65*

Avail C:N 0.19 0.45 -0.21 -0.14 0.63* -0.35 -0.18 -0.14 -0.17

TN -0.43 -0.10 -0.20 0.07 0.22 -0.03 0.39 -0.14 0.64*

C:N 0.21 0.23 -0.03 0.13 -0.17 0.17 0.31 0.32 -0.05

Fe3+ -0.13 -0.10 -0.01 -0.17 -0.30 -0.03 -0.04 -0.32 0.16

Cu 0.23 -0.02 0.16 -0.24 0.006 -0.20 0.47 0.10 0.05

AP

PE

ND

IX D

AT

A C

HA

PT

ER

4 C

ON

TIN

UE

D

Page 116: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

106

106

Appendix Table 4.12a Pearson Correlation ( r ) between qnorB and nosZ abundances and soil parameter by month under prairie management

N= 9 May August October

Parameters qnorB nosZ qnorB nosZ qnorB nosZ

qnorB - 0.32 - 0.37 - 0.73**

nosZ 0.32 - 0.37 - 0.73** -

WFPS 0.25 0.31 0.44 0.37 0.15 -0.08

pH -0.16 0.06 -0.39 0.18 -0.01 0.05

EOC -0.10 0.21 0.09 0.51 -0.09 -0.03

TC -0.15 0.34 0.14 0.01 0.02 0.18

NH4+ 0.08 0.72** 0.30 0.14 0.13 0.07

NO3-

-0.24 0.12 0.36 0.12 -0.25 -0.02

TIN -0.21 0.35 0.11 -0.01 0.07 0.19

Avail C:N -0.18 -0.11 -0.36 0.36 -0.23 -0.13

TN -0.22 0.33 -0.29 -0.43 0.08 0.20

C:N 0.35 0.24 0.05 -0.13 -0.07 -0.10

Fe3+

0.50 -0.08 0.33 -0.24 0.24 -0.11

Cu 0.01 -0.25 -0.09 -0.22 -0.11 -0.44

Appendix Table 4.12b Pearson Correlation ( r ) between qnorB, nosZ, qnorB:16S, qnosZ:16S , and qnor:nosZ abundances and soil parameter

by month under prairie management

N= 9 May August October

Parameters qnorB:16S nosZ:16S qnorB:nosZ qnorB:16S nosZ:16S qnorB:nosZ qnorB:16S nosZ:16S qnorB:nosZ

WFPS 0.24 0.49 0.18 0.19 0.31 0.07 0.45 0.23 0.30

pH 0.35 0.55 0.17 -0.52 0.20 -0.51 -0.08 0.05 -0.09

EOC 0.35 0.70* 0.28 -0.30 0.35 -0.37 0.09 0.35 -0.07

TC 0.07 0.16 0.08 0.12 -0.01 0.11 0.06 0.67* -0.23

NH4+ -0.13 0.25 0.29 0.14 -0.05 0.14 0.21 0.30 0.06

NO3-

0.18 0.16 0.005 0.36 0.29 0.21 -0.20 0.22 -0.28

TIN 0.07 0.10 0.03 0.12 0.002 0.10 -0.20 0.48 -0.18

Avail C:N 0.17 0.54 0.13 -0.68* 0.13 -0.63* 0.04 0.20 -0.12

TN 0.05 0.07 0.03 0.23 0.30 0.10 -0.03 0.48 -0.18

C:N 0.16 0.51 0.27 0.19 0.02 0.16 0.16 0.22 0.05

Fe3+

-0.23 -0.23 0.04 0.56 -0.07 0.51 0.23 -0.58* 0.46

Cu 0.09 0.10 -0.20 0.01 -0.34 0.12 0.30 -0.48 0.48

AP

PE

ND

IX D

AT

A C

HA

PT

ER

4 C

ON

TIN

UE

D

Page 117: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

107

107

Appendix Table 4.13a Pearson Correlation ( r ) between qnorB, nosZ, qnorB:16S, qnosZ:16S , and qnor:nosZ abundances

and soil parameter under soil types by month

May

N= 6 Clarion Webster

Parameters qnorB nosZ qnorB:16S nosZ:16S qnorB:nosZ qnorB nosZ qnorB:16S nosZ:16S qnorB:nosZ

qnorB - 0.69 - - - 0.61 - - - -

nosZ 0.69 - - - - - 0.61 - - -

WFPS -0.03 0.53 -0.41 0.18 -0.48 0.23 0.61 -0.39 0.27 -0.17

pH -0.20 -0.02 -0.34 -0.54 -0.27 -0.41 0.17 -0.27 0.74 0.38

EOC -0.07 0.63 -0.64 -0.49 -0.61 -0.35 0.36 -0.79 0.62 0.24

TC -0.05 0.64 -0.65 -0.57 -0.60 -0.02 0.69 -0.59 0.68 -0.43

NH4+ -0.07 0.26 -0.33 -0.30 -0.31 0.03 -0.24 0.01 -0.39 -0.43

NO3-

-0.36 -0.41 -0.03 0.63 -0.16 -0.04 0.04 -0.14 0.04 0.52

TIN -0.16 0.55 -0.72 -0.59 -0.68 -0.02 0.72 -0.58 0.73 -0.54

Avail C:N 0.05 0.01 0.04 -0.10 0.06 -0.43 0.30 -0.84* 0.63 -0.03

TN -0.13 0.59 -0.70 -0.57 -0.65 -0.04 0.70 -0.59 0.73 -0.55

C:N 0.86* 0.54 0.68 0.02 0.73 0.17 0.39 -0.44 0.04 0.22

Fe3+

0.68 0.44 0.63 0.69 0.58 -0.01 0.17 -0.52 -0.11 -0.90*

Cu 0.09 -0.28 0.44 0.73 0.34 -0.12 -0.18 0.17 0.02 -0.81

Canisteo Okoboji

qnorB nosZ qnorB:16S nosZ:16S qnorB:nosZ qnorB nosZ qnorB:16S nosZ:16S qnorB:nosZ

qnorB - 0.35 - - - - 0.20 - - -

nosZ 0.35 - - - - 0.20 - - - -

WFPS 0.15 0.39 -0.08 -0.02 -0.04 0.52 0.77 -0.14 0.15 -0.36

pH -0.41 0.19 -0.77 0.12 -0.52 0.14 0.31 0.53 -0.09 -0.66

EOC -0.17 0.29 -0.77 -0.15 -0.32 0.27 0.44 0.61 0.18 0.80

TC 0.13 0.38 -0.53 -0.33 -0.04 -0.08 0.71 -0.60 0.07 -0.75

NH4+ -0.07 0.40 -0.70 -0.17 -0.27 0.33 0.80 -0.17 0.46 0.29

NO3-

0.65 0.22 0.85* -0.12 0.58 -0.49 -0.88* 0.47 -0.33 -0.09

TIN 0.21 0.34 -0.45 -0.40 0.06 -0.19 0.74 -0.64 0.18 -0.79

Avail C:N -0.58 0.01 -0.93** 0.12 -0.62 0.40 0.71 0.33 0.34 -0.83*

TN 0.21 0.34 -0.46 -0.40 0.05 -0.21 0.72 -0.67 0.16 -0.79

C:N -0.56 0.21 -0.36 0.64 -0.70 0.52 0.39 -0.08 -0.32 -0.22

Fe3+

0.67 -0.08 0.08 -0.91* 0.75 0.51 0.44 -0.79 0.47 -0.18

Cu -0.30 -0.68 -0.25 -0.20 0.01 -0.46 0.36 -0.86 0.36 0.05

AP

PE

ND

IX D

AT

A C

HA

PT

ER

4 C

ON

TIN

UE

D

Page 118: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

108

108

Appendix Table 4.13b Pearson Correlation ( r ) between qnorB, nosZ, qnorB:16S, qnosZ:16S , and qnor:nosZ abundances

and soil parameter under soil types by month continued

August

N= 6 Clarion Webster

Parameters qnorB nosZ qnorB:16S nosZ:16S qnorB:nosZ qnorB nosZ qnorB:16S nosZ:16S qnorB:nosZ

qnorB - 0.30 - - - - 0.58 - - -

nosZ 0.30 - - - - 0.58 - - - -

WFPS 0.17 0.37 -0.23 0.11 -0.21 -0.18 -0.49 0.12 -0.52 0.32

pH -0.19 -0.30 0.006 -0.38 0.10 0.12 0.34 0.04 0.53 -0.22

EOC 0.90* 0.31 0.47 -0.10 0.40 -0.64 -0.17 -0.55 0.39 -0.55

TC 0.88* 0.33 0.44 0.09 0.33 -0.87* -0.23 -0.29 -0.23 -0.10

NH4+ 0.62 0.18 0.36 -0.04 0.30 -0.62 -0.34 -0.45 0.08 -0.35

NO3-

-0.46 -0.56 0.08 -0.42 0.17 0.89* 0.61 0.44 -0.13 0.36

TIN 0.82* 0.42 0.36 0.25 0.23 -0.46 -0.29 -0.48 -0.04 -0.31

Avail C:N 0.88* 0.54 0.25 0.11 0.17 -0.87* -0.17 -0.79 0.60 -0.81

TN 0.84* 0.37 0.42 0.19 0.29 -0.88* -0.39 -0.78 0.11 -0.59

C:N 0.45 0.03 0.30 -0.29 0.32 0.42 -0.23 0.54 -0.74 0.71

Fe3+

-0.80 -0.20 -0.49 0.12 -0.42 0.04 0.12 -0.14 -0.07 -0.07

Cu -0.56 0.08 -0.56 0.18 -0.49 0.64 -0.12 0.91* -0.50 0.85

Canisteo Okoboji

qnorB nosZ qnorB:16S nosZ:16S qnorB:nosZ qnorB nosZ qnorB:16S nosZ:16S qnorB:nosZ

qnorB - 0.30 - - - - 0.45 - - -

nosZ 0.30 - - - - 0.45 - - - -

WFPS -0.53 -0.53 -0.28 0.18 -0.29 0.12 0.48 -0.27 0.43 -0.32

pH 0.58 0.55 -0.002 0.26 -0.08 -0.28 0.07 -0.42 -0.38 -0.34

EOC -0.48 0.27 -0.49 0.12 -0.45 0.54 0.69 -0.11 -0.04 -0.10

TC -0.64 -0.23 -0.31 -0.24 -0.17 0.30 0.29 0.09 0.29 0.04

NH4+ -0.39 0.15 -0.3 -0.04 -0.31 0.73 0.54 0.28 0.25 0.23

NO3-

-0.01 0.65 -0.21 0.89* -0.47 0.54 0.41 0.25 0.50 0.16

TIN -0.62 -0.19 -0.31 -0.19 -0.19 0.38 0.28 0.16 0.25 0.12

Avail C:N -0.48 0.14 -0.56 -0.33 -0.35 -0.48 0.03 -0.59 -0.34 -0.51

TN -0.62 -0.18 -0.32 -0.20 -0.20 -0.02 -0.51 0.37 0.37 0.27

C:N 0.26 -0.17 0.25 -0.15 0.26 -0.27 -0.12 -0.07 0.61 -0.16

Fe3+

-0.01 -0.72 0.54 -0.04 0.46 -0.02 -0.49 0.59 0.30 0.43

Cu 0.72 0.35 0.15 -0.03 0.13 -0.24 0.08 -0.32 -0.0003 -0.31

AP

PE

ND

IX D

AT

A C

HA

PT

ER

4 C

ON

TIN

UE

D

Page 119: Biogeography of denitrifying bacteria in a prairie and agro-ecosystem

109

109

Appendix Table 4.13c Pearson Correlation ( r ) between qnorB, nosZ, qnorB:16S, qnosZ:16S , and qnor:nosZ abundances

and soil parameter under soil types by month continued

October

N= 9 Clarion Webster

Parameters qnorB nosZ qnorB:16S nosZ:16S qnorB:nosZ qnorB nosZ qnorB:16S nosZ:16S qnorB:nosZ

qnorB - 0.30 - - - - 0.64 - - -

nosZ 0.30 - - - - 0.64 - - - -

WFPS 0.85* 0.09 0.74 -0.59 0.86 0.26 0.30 0.22 0.55 0.03

pH -0.56 -0.41 -0.63 -0.08 -0.50 -0.02 0.55 -0.46 -0.01 -0.59

EOC -0.25 -0.04 -0.07 -0.77 -0.25 -0.23 -0.36 -0.10 -0.44 0.07

TC 0.36 0.23 0.29 -0.90* 0.32 -0.07 -0.02 0.06 0.34 -0.07

NH4+ 0.12 0.005 0.58 -0.89* 0.12 -0.22 -0.02 -0.20 0.03 -0.26

NO3-

-0.36 -0.20 -0.25 0.94* -0.33 0.33 0.39 -0.01 -0.12 0.04

TIN 0.40 0.19 0.34 -0.90* 0.38 -0.09 -0.06 0.08 0.36 -0.05

Avail C:N -0.30 -0.05 -0.28 -0.68 -0.30 -0.36 -0.55 -0.18 -0.72 0.09

TN 0.37 0.21 0.32 -0.88* 0.34 -0.08 -0.02 0.06 0.37 -0.08

C:N 0.004 0.20 -0.09 -0.41 -0.04 0.007 -0.08 -0.01 -0.26 0.09

Fe3+

0.27 0.41 0.47 0.20 0.20 0.02 -0.59 0.43 -0.19 0.62

Cu 0.30 -0.56 0.78 0.10 0.42 0.17 -0.35 0.49 0.11 0.57

Canisteo Okoboji

qnorB nosZ qnorB:16S nosZ:16S qnorB:nosZ qnorB nosZ qnorB:16S nosZ:16S qnorB:nosZ

qnorB - 0.49 - - - - 0.74 - - -

nosZ 0.49 - - - - 0.74 - - - -

WFPS -0.04 -0.06 -0.56 -0.78 0.03 0.03 0.03 -0.59 -0.22 -0.31

pH 0.09 0.50 0.005 0.48 -0.35 -0.47 -0.47 0.14 -0.14 -0.42

EOC 0.42 0.17 -0.41 -0.92** 0.27 -0.39 -0.43 -0.29 0.77 0.03

TC -0.07 0.39 -0.87* -0.56 -0.44 0.11 0.44 -0.51 0.04 -0.42

NH4+ 0.30 -0.07 -0.25 -0.85 0.38 0.48 0.36 0.11 -0.13 0.20

NO3-

-0.57 -0.65 0.35 0.44 0.02 -0.57 -0.72 0.07 0.48 0.17

TIN -0.07 0.35 -0.88* -0.61 -0.41 0.12 0.41 -0.52 -0.03 -0.39

Avail C:N 0.25 0.45 -0.44 -0.35 -0.17 -0.74 -0.69 -0.51 0.99** -0.13

TN -0.10 0.33 -0.88* -0.60 -0.42 0.13 0.43 -0.49 -0.02 -0.39

C:N 0.34 0.65 0.12 0.52 -0.26 -0.47 -0.02 -0.33 0.97* -0.65

Fe3+

-0.14 -0.42 0.43 0.18 0.29 0.02 -0.30 0.18 -0.65 0.44

Cu 0.40 -0.27 0.52 -0.22 0.67 -0.51 -0.36 -0.12 0.23 -0.24

AP

PE

ND

IX D

AT

A C

HA

PT

ER

4 C

ON

TIN

UE

D