FEMS Microbiology Ecology Advance Access published December 8, 2014

1

Title: Ammonia-oxidizing archaea respond positively to inorganic nitrogen addition in desert

2

soils

3

Authors: Yevgeniy Marusenko1*, Ferran Garcia-Pichel1, Sharon J. Hall1

4 5

Affiliation: 1School of Life Sciences, Arizona State University, Tempe, AZ, 85287

6 7

Correspondence: *Yevgeniy Marusenko, 602-703-7984, [email protected]

8

Current address: 3312 Biological Sciences III,

9

University of California, Irvine, CA, 92697, USA

10 11

11

Abstract

12

In soils, nitrogen (N) addition typically enhances ammonia oxidation (AO) rates and increases

13

the population density of ammonia-oxidizing bacteria (AOB), but not that of ammonia-oxidizing

14

archaea (AOA). We asked if long-term inorganic N addition also has similar consequences in

15

arid land soils, an understudied yet spatially ubiquitous ecosystem type. Using Sonoran Desert

16

top soils from between and under shrubs within a long-term N-enrichment experiment, we

17

determined community concentration-response kinetics of AO and measured the total and

18

relative abundance of AOA and AOB based on amoA gene abundance. As expected, N addition

19

increased maximum AO rates and the abundance of bacterial amoA genes compared to the

20

controls. Surprisingly, N addition also increased the abundance of archaeal amoA genes. We did

21

not detect any major effects of N addition on ammonia-oxidizing community composition. The

22

ammonia-oxidizing communities in these desert soils were dominated by AOA as expected (78%

23

of amoA gene copies were related to Nitrososphaera), but contained unusually high contributions

24

of Nitrosomonas (18%) and unusually low numbers of Nitrosospira (2%). This study highlights

25

unique traits of ammonia-oxidizers in arid lands, which should be considered globally in

26

predictions of AO responses to changes in N availability.

27 28

2

28

Introduction

29 30

Since the early and influential work of Sergei Winogradsky (1890), bacteria were thought to be

31

the only biological agents of ammonia oxidation (AO). However, the deployment of molecular

32

detection techniques in the last three decades has revealed that Thaumarchaeota in the Archaea

33

domain contribute to AO as well (Konneke et al., 2005). High-throughput sequencing and

34

molecular-fingerprinting studies show the presence of genes attributable to diverse groups of

35

ammonia-oxidizing archaea (AOA) and bacteria (AOB) in a wide variety of environments

36

(Purkhold et al., 2000; Leininger et al., 2006; Prosser & Nicol, 2008; Pester et al., 2012). Even

37

though AOA outnumber AOB in many ecosystems (Leininger et al., 2006; Adair & Schwartz,

38

2008; Wessen et al., 2010), this dominance does not always equate to AOA contributing to AO

39

more than AOB (Jia & Conrad, 2009; Di et al., 2009; Adair & Schwartz, 2011). It remains

40

unclear why the abundance of AOA is often unrelated to AO rates (Shen et al., 2008; Wessen et

41

al., 2010). AO fluxes may depend not only on population size, but also on community

42

composition due to differential substrate affinities and ecophysiological sensitivities among and

43

within the AOA and AOB (Kowalchuk & Stephen, 2001; Bollmann et al., 2002; Schleper &

44

Nicol, 2010).

45 46

A review of literature reveals that mixed ammonia-oxidizer communities are often dominated by

47

one particular phylotype (Prosser, 1989; Kowalchuk & Stephen, 2001; Zhalnina et al., 2012; He

48

et al., 2012). However, it is uncertain if and how this outcome is determined by environmental

49

properties. For instance, while culture work shows that Nitrosomonas strains (AOB) prefer

50

ammonia-rich conditions (Taylor & Bottomley, 2006), Nitrosospira-related clusters (AOB)

51

commonly outnumber Nitrosomonas spp. in fertilized soils and also in low-NH4+, pristine soils 3

52

(Jordan et al., 2005; Chu et al., 2007). Additionally, AOB are preferentially enriched after

53

inorganic nitrogen (N) fertilization in the ecosystems studied to date – such as in relatively low-

54

pH soils that receive high rates of precipitation or water inputs – while AOA may respond

55

positively only in cases when NH3/NH4+ is supplied through organic matter mineralization (Offre

56

et al., 2009; He et al., 2012; Hatzenpichler, 2012; Levicnik-Hofferle et al., 2012). These

57

examples suggest that indeed changes in N availability such as through N deposition or

58

fertilization may control AO rates in soils through community compositional shifts (Avrahami &

59

Bohannan, 2007; Tourna et al., 2010; Prosser & Nicol, 2012).

60 61

Arid environments are vastly underrepresented in the AO research literature (Johnson et al.,

62

2005; Marusenko et al., 2013b; Sher et al., 2013), but there is reason to believe that arid lands

63

may harbor populations with different adaptations compared to the more studied temperate soils.

64

For example, arid soils are exposed to prolonged drought and rapid pulses of precipitation and

65

nutrients (Schimel et al., 2007; Collins et al., 2008), which require complex and fast genetic

66

regulation from soil microbes (Rajeev et al., 2013). Furthermore, arid soils are often alkaline and

67

can reach up to 50°C in the summer. They are typically dry with low organic matter content and

68

low N mineralization rates especially in non-vegetated areas between shrubs (Austin et al.,

69

2004; Schade & Hobbie, 2005), which may select for the most oligotrophic of ammonia-

70

oxidizers. These ubiquitous soils also experience intensive management, including watering and

71

fertilizer inputs, both in agricultural and urban residential areas (Warren et al., 1996; Davies &

72

Hall, 2010). As a result, anthropogenic activities and atmospheric deposition are altering

73

resource availability and the N cycle in soils of water-limited environments (McCrackin et al.,

74

2008; Hall et al., 2009; Hall et al., 2011; Marusenko et al., 2013a).

4

75 76

Here we tested the effect of long-term inorganic N addition on AO processes and ammonia-

77

oxidizing microorganisms (AOM) in arid land soils, assessing the AO kinetics in bulk soil and

78

characterizing AOA and AOB by sequencing the environmental amoA gene, which encodes a

79

subunit of the ammonia-monooxygenase (AMO) enzyme. We hypothesized that N addition

80

would cause the absolute and relative abundance of ammonia-oxidizers to shift from AOA-

81

dominated in oligotrophic native (unfertilized) soils to AOB-dominated in NH4+-rich conditions,

82

as has been found in other soils. Consequently, this population replacement would enhance

83

overall AO rates and cell-specific AO rates, but decrease affinity between the enzyme and the

84

substrate. Using common patch types in arid lands, we further expected that the decline of AOA

85

relative to AOB under N addition would be less dramatic in relatively fertile soils under shrubs

86

than in areas away from plants.

87 88

Materials and Methods

89 90

Study area description

91 92

Our site is in the northern Sonoran Desert at ~620 m elevation in Lost Dutchman State Park, AZ,

93

USA (coordinates: N 33.459372 S -111.484956), located east of the Phoenix metropolitan area

94

and within boundaries of the Central Arizona–Phoenix Long-Term Ecological Research area

95

(http://caplter.asu.edu). Soils are classified as Typic Haplargids, a subgroup of Aridisols. We

96

measured soil AO rates and community parameters from two randomly assigned 20 m x 20 m

97

plots, one that received N fertilizer as NH4NO3 (applied as solid by hand biannually at 60 kg N

5

98

ha-1 yr-1 from 2005-2012) and another that served as an unfertilized control (see Hall et al., 2011

99

for further description about the plots). Nitrogen deposition in this area is 7.3 kg N ha-1 yr-1

100

(Cook, 2014). Plant cover (~60%) within our study plots is dominated by the native shrubs

101

creosote bush (Larrea tridentata [DC.] Coville), bursage (Ambrosia spp.). Plots did not contain

102

any N-fixing trees. Mean annual temperature is 22.3°C, with the coldest and warmest months

103

averaging 3.7°C and 41.9°C, respectively (2005-2012; NCDC, 2013). Mean annual precipitation

104

is 272 mm but is highly variable year to year. Rainfall is bimodally distributed between summer

105

monsoon events and low-intensity winter storms (WRCC, 1985).

106 107

Sample collection

108 109

Surface soil samples were collected in late January of 2012, one month after winter storms. In

110

each of the control and N addition plots, three soil samples were collected from each of two

111

patch types to explore N treatment effects in typical desert environments: between plants

112

(hereafter called 'interplant') and under canopies of the common shrub L. tridentata ('under

113

plant'). Mature/dark soil biological crusts were low in abundance within the plot area and were

114

avoided for sampling. Early colonization by biocrust organisms is widespread in the region

115

(Rosentreter et al., 2007) but is not yet formed to visibility at our site locations. Each soil sample

116

consisted of two 0-5 cm (depth) x 7 cm (diameter) cores taken 5 cm apart. In total, we collected

117

twelve soil samples consisting of three replicate soil samples from each plot (treatment, control)

118

and patch type (interplant, under plant) (3x2x2 = 12 samples). Soil samples were processed

119

independently for all analyses (soil properties, AO rates, quantitative PCR, pyrosequencing).

120

6

121

Laboratory methods and soil properties

122 123

Following collection, samples were transported on ice to the lab, sieved to < 2 mm, and

124

homogenized. Soils were at 3-5% soil moisture upon collection and were analyzed within 24 h

125

for all soil properties and processes. Two subsamples (2 g each) from each homogenized soil

126

sample were frozen in liquid N and stored at -80oC until DNA extraction within one month.

127

Duplicate DNA extracts were combined prior to molecular processing methods to obtain one

128

determination per sample.

129 130

Soils were processed for pH (1:2 soil to DI H2O), water holding capacity (% WHC;

131

gravimetrically), organic matter content (% SOM; loss on ignition), and extractable NH4+, nitrite

132

(NO2-), and nitrate (NO3-) content (2M KCl extraction, colorimetric analysis), following standard

133

methods (Sparks et al., 1996; Marusenko et al., 2013a). Data reported for each of the three field

134

replicates is an average of laboratory triplicates.

135 136

Ammonia oxidation rates using the shaken-slurry assay

137 138

In situ net rates of potential AO were measured under various levels of N addition (see

139

“Ammonia oxidation kinetics” below) using the shaken-slurry method (hereafter as “slurry AO

140

rates”), in which oxygen and substrate diffusion is not limiting (Hart et al., 1994; Norton & Stark

141

2011). The direct product of AO was measured as NO2- accumulation after inclusion of chlorate

142

(NaClO3), a NO2--oxidation inhibitor (Belser & Mays 1980). Using NO3 - as a proxy for AO was

143

unsuitable since NO2- build-up is common in natural dryland conditions (Gelfand & Yakir 2008).

7

144

The shaken-slurry assays contained 10 g soil in 100 mL solution of 0.015 mol·L -1 NaClO3, and

145

0.2 mol·L-1 K2HPO4 and 0.2 mol·L-1 KH2PO4 to buffer pH at 7.2. Slurries and no-soil blanks

146

were continuously aerated in solution by mixing at 180 rpm on a reciprocal shaker in the dark.

147

Homogenized slurry aliquots were removed at four time points over 6 h and amended with

148

several drops of MgCl2 + CaCl2 (0.6 M) to flocculate soil particles. Aliquots were then

149

centrifuged at 3000 × g and supernatant was filtered through pre-leached Whatman #42 ashless

150

filters. The supernatants were stored at 4°C and analyzed within 24 h. Net rates of slurry AO

151

were calculated as the linear increase in NO2- content from 0 to 6 h, measured colorimetrically

152

using a Lachat Quikchem 8000 autoanalyzer. Consistent with literature showing that metabolism

153

of ammonia oxidizers can be activated and responsive to the environment at the scale of hours,

154

especially for AOB (Wilhelm et al., 1998; Placella & Firestone 2013), NO2- accumulation in our

155

assays was linear from 0 to 6 h.

156 157

Ammonia oxidation rates in static incubation

158

As a secondary method to slurry AO rates, we also measured AO following various levels of N

159

addition in a modified method using NaClO3 inhibition in static, 48 h aerobic incubations of bulk

160

soil (“static AO rates” from here on; Nishio & Fujimoto 1990; Hart et al., 1994; Low et al.,

161

1997). Although substrate diffusion may be limited in aerobic incubations to fully quantify

162

enzyme activity, we used this method to independently assess AO in conditions more

163

representative of the upland desert environment compared to the shaken-slurry assays, which

164

assesses aerated AO potential. Ten g of soil were brought to 60% WHC using water and NaClO3

165

(15 mM) in plastic cups. Soil in one cup was extracted at the onset and a second cup extracted

166

after incubation for 2 days in the dark. Soils were extracted in 50 mL of 2 M KCl followed by

8

167

shaking for 1 h and filtering through pre-leached Whatman #42 ashless filters. The extracts were

168

stored at 4°C and analyzed colorimetrically within 24 h. Net rates of static AO were calculated

169

as the increase in NO2- content between 0 and 48 h.

170 171

Ammonia oxidation kinetics

172 173

The relationship between substrate availability and reaction rate can be measured to discern

174

functional parameters of microbial communities. Although this is not a true enzyme kinetics

175

study, we applied similar calculations to our concentration-response kinetics approach to model

176

relationships for AO (slurry and static) in bulk soils based on the Michaelis-Menten equation

177

(Martens-Habbena & Stahl 2011; Prosser & Nicol 2012), as has been successfully applied to soil

178

assays (Koper et al. 2010):

179

V = (Vmax × S) / (Km + S)

180

In this equation, the NH4+ concentration (S) and AO rate (V) are used to estimate the maximum

181

AO rate (Vmax) and half-saturation constant (Km; inverse of enzyme and substrate affinity). To

182

estimate AO kinetics under oligotrophic conditions in the shaken-slurry assay, we removed pre-

183

existing NH4+ from soils to obtain the least variable and lowest residual substrate availability

184

(Widmer et al., 1989; Koper et al., 2010; Norton & Stark 2011). Prior to the shaken-slurry assay,

185

5 g soil was mixed in 45 mL of potassium phosphate solution and centrifuged at 3200 × g for 1

186

min before discarding the N-containing supernatant. The resulting soil pellet from two

187

preparations was combined to compose 10 g total soil from each plot (treatment, control), patch

188

type (interplant, under plant), and soil sample replicate (x 3). Inorganic N was then supplemented

189

as (NH4)2SO4 mixed with DI water to eight final concentrations in the slurry ranging from 0-22.5

9

190

mM. In total we evaluated AO rates using 96 different soil preparations (12 soil samples x 8

191

NH4+ concentrations) per method (shaken-slurry assay, static incubation). In the static

192

incubations, we excluded the N removal step as to minimize soil disturbance. Soils were

193

supplemented with (NH4)2SO4 in solution to produce final concentrations ranging from 0-50 µg

194

NH4+-N g-1 (0-22.5 mM). The 0 µg NH4+-N g-1 addition (only includes pre-existing NH4+) was

195

used to estimate background net rates of AO. As a rough indicator of the N addition effect on the

196

relative importance of NH4+ mineralization and nitrification, we also measured the net rate of

197

NH4+ gain (production processes dominate) and loss (consumption processes dominate) during

198

the static incubation experiment. In the assay, some of the NH4+ consumption processes are

199

likely minimized due to sieving of soil (exclusion of large NH4+-assimilating plant roots) and

200

lower laboratory temperature compared to natural conditions (reduced volatilization).

201 202

DNA extraction and purification

203 204

DNA was extracted using three freeze-thaw cycles followed by 30 min incubation at 50°C with

205

proteinase K and silica bead beating for chemical and mechanical cell lysis (Garcia-Pichel et al.,

206

2001). The lysate was purified by phenol:chloroform:isoamyl alcohol (25:24:1) extraction,

207

followed by DNA precipitation in 100% ethanol for 12 h at -80°C. DNA concentration and

208

quality was assessed on an agarose gel stained in ethidium bromide and imaged using a Fluor-S

209

Multi-Imager (BioRad Laboratories, CA, USA) with an EZ Load Precision Molecular Mass

210

Standard (BioRad). Bands of DNA were excised from a low-melt agarose gel, homogenized with

211

a tip in a microcentrifuge tube, allowed to diffuse out into sterile H2O for 12 h, and followed by

212

15 min centrifugation to collect DNA in the supernatant.

10

213 214

Quantitative PCR

215 216

DNA was used for quantitative PCR (qPCR) with the following amoA primers: CrenamoA616r

217

(GCCATCCABCKRTANGTCCA; Tourna et al., 2008) and CrenamoA23f for the AOA

218

(ATGGTCTGGCTWAGACG); and amoA1f mod (GGGGHTTYTACTGGTGGT; Stephen et

219

al., 1999) and AmoA-2R’ for the AOB (CCTCKGSAAAGCCTTCTTC; Okano et al., 2004;

220

Junier et al., 2008). qPCR reactions contained 10 µL iTaq SYBRGreen Master Mix (BioRad),

221

250 nM final concentration of each primer (AOA or AOB), 1 ng of environmental DNA, and

222

molecular grade H2O to bring each reaction to a final volume of 20 µL. The reaction conditions

223

were as follows: initial denaturation for 150 s at 95°C followed by 45 cycles of 15 s at 95°C, 30 s

224

at 55°C, and 30 s at 72°C, and a final dissociation step to obtain the melting curve at 95°C, 60°C,

225

and 95°C for 15 s each. Standard curves were generated using templates from Nitrosomonas

226

europaea ATCC 19718 (bacterial amoA; R2 = 0.99) and a putative AOA clone (archaeal amoA;

227

R2 = 0.99) for a dilution series spanning 102-1010 gene copies per reaction. Melting curves were

228

checked to verify the quality of each reaction, and to ensure the absence of primer-dimers. We

229

report only determinations for which Ct values could be interpolated within our standard curves.

230

Each amoA abundance value (number of gene copies) reported is an average of analytical

231

triplicate qPCR reactions of the same DNA extract.

232 233

Pyrosequencing

234

11

235

Purified DNA extracts were shipped to a commercial laboratory for standard PCR and bTEFAP

236

pyrosequencing (Dowd et al., 2008). Commercial primers for PCR were amoA-1F

237

(GGGGTTTCTACTGGTGGT; Rotthauwe et al., 1997) and amoA-2R for AOB

238

(CCCCTCKGSAAAGCCTTCTTC), and Arch-amoAF (STAATGGTCTGGCTTAGACG;

239

Francis et al., 2005) and Arch-amoAR for AOA (GCGGCCATCCATCTGTATGT) used with a

240

HotStarTaq Plus Master Mix Kit (Qiagen, CA, USA). PCR conditions were as follows: 180 s at

241

94°C followed by 28 cycles of 30 s at 94°C, 40 s at 53°C, and 60 s at 72°C, and final elongation

242

for 5 min at 72°C. PCR amplicons were mixed in equal concentrations and purified using

243

Agencourt Ampure beads (Agencourt Bioscience Corporation, MA, USA). Sequencing utilized

244

Roche 454 FLX titanium instruments and reagents.

245 246

Bioinformatics and phylogenetic analyses of amoA

247 248

Pyrosequencing data were processed and analysed in Qiime (Caporaso et al., 2010b), with

249

necessary pipeline adjustments to process functional gene data (i.e., amoA) as described in detail

250

in the notes and script file (http://www.yevmarusenko.com/research/Marusenko_Qiime.txt).

251

Sequences (452 bp long) were clustered into operational taxonomic units (OTUs) using UClust

252

(Edgar, 2010). Representative sequences (one per OTU) were aligned with Pynast (Caporaso et

253

al., 2010a). Based on nomenclature classification for AOA in Pester et al. (2012) and for AOB in

254

Koops et al. (2006), a taxonomic assignment was made for each OTU using a template reference

255

database created from sequences of known pure isolates, enrichments, and other characterized

256

AOA and AOB from previous studies (reference database available at website mentioned above).

257

Groups of sequences were clustered at 97% nucleotide similarity to be inclusive of OTUs at fine

12

258

levels of resolution for phylogenetic and statistical analyses that otherwise may be missed at

259

lower identity thresholds. For AOA, we excluded one replicate each in the interplant control and

260

N addition samples because of their poor quality of the pyrosequencing data. The minimum

261

number of high-quality sequences after filtering was 525 for AOA and 950 for AOB, with

262

sufficient rarefied analysis producing 179 OTUs for AOA and 325 OTUs for AOB (Total

263

number of sequences >200bp prior to quality filtering: AOA, 9,830; AOB, 29,569). Phylogenetic

264

analyses were carried out on a single alignment file (separately for AOA and AOB) that included

265

sequences from our Qiime pipeline, as well as the reference sequences described above. All

266

sequences were combined and realigned using default parameters for Muscle and analyzed by

267

the tree-building module of the MEGA 5 software with the following parameters: Neighbor-

268

joining method, Jukes-Cantor nucleotide substitution model, 100 bootstrap replicates, uniform

269

rates among sites, and pairwise gap-data deletion (Tamura et al., 2011). Raw sequence reads for

270

the entire project have been deposited in the Sequence Read Archive (SRA) at NCBI with

271

accession number SRSRX738968 for the AOA data and SRX739281 for the AOB data.

272 273

Statistics

274 275

Statistical tests were carried out using Qiime for α and β diversity measures on processed

276

pyrosequencing data, while all other analyses were in SPSS (v20.0 Windows). All soil

277

properties, AO rates, and amoA abundance data were tested for linear model assumptions in

278

SPSS using normal probability plots (for normality) and Levene’s test (for equal variance), and

279

transformed (natural log) when necessary. Individual two-way analysis of variance (ANOVA)

280

tests were used to evaluate the effects of plants (‘Patch’) and N addition (‘Treatment’) on the

13

281

following dependent variables: amoA gene abundance (per g soil and per ng extractable DNA,

282

separately for AOA and AOB), AOA to AOB ratio, slurry Vmax AO rates, static Vmax AO rates,

283

amoA-copy specific AO rates, net NH4+ change (averaged across supplemented NH4+

284

concentrations), and each of the soil properties. Significant interactions between patch and

285

treatment were evaluated further using one-way ANOVA (α = 0.025). The copy specific rates

286

were calculated using Vmax AO rate (slurry and static) divided by the number of amoA gene

287

copies per g soil. We used bivariate Pearson correlations to assess relationships between soil

288

properties vs. community parameters (amoA data and AO rates) across all samples. We used

289

linear regression analyses to assess relationships between amoA gene abundance at the domain

290

level vs. Vmax AO rates and also analyzed amoA gene abundance of individual OTUs vs. soil

291

properties and AO rates. In Qiime, we tested for the effect of N addition on OTU-based

292

communities separately for AOA and AOB per patch type, using only strictly relevant diversity

293

metrics (Lozupone & Knight, 2005; Caporaso et al., 2010b): α diversity (Shannon’s diversity,

294

observed richness, and phylogenetic diversity [PD]) and β diversity (weighted and unweighted

295

Unifrac, the multivariate group dispersion analogue of Levene's test [PERMDISP], and analysis

296

of similarity [ANOSIM]).

297 298

Results

299 300

Effects of N addition on the abundance of amoA genes and the kinetics of ammonia oxidation

301 302

To aid in interpretation of long-term N addition effects on amoA gene abundance and AO rates,

303

we considered the influence of soil properties and the relative importance of fertilizer N vs.

14

304

ammonification as a possible NH4+ source. Long-term N addition clearly resulted in an

305

accumulation of NO2 -, NO3-, and NH4+, regardless of patch type (Table 1). Also as expected, soil

306

organic matter was higher in soil under plants than between plants. N addition slightly acidified

307

these alkaline soils – an effect known to worsen conditions for AO (Arp & Stein, 2003) – and yet

308

AO rates still increase in these N-amended desert plots (Suppl. Table 1). Both types of AO rates

309

we measured (maximum static and slurry AO rates) were strongly predicted by pH (negative

310

correlation) and soil organic matter (positive correlation). Background AO rates measured in

311

unamended incubations, however, were most strongly and positively related to NH4+

312

concentration across all patch types and N treatments. Additionally, N addition significantly

313

increased net rates of NH4+ loss in both interplant and under plant patch types (Patch, P = 0.56;

314

Treatment, P < 0.01). These data suggest that N addition stimulated NH4+ loss from consumption

315

processes (e.g., NH3/NH4+ oxidation, microbial immobilization) relatively more than it increased

316

NH4+ concentrations from organic N mineralization.

317 318

Surprisingly, N addition increased abundance of archaeal amoA genes compared to controls,

319

regardless of the measure used (copies per g-1 soil, Fig. 1; or copies per total community DNA,

320

904 vs 548 archaeal amoA copies/ng DNA; in soils between plants). As expected from many

321

other studies, long-term N addition also increased the abundance of bacterial amoA gene copies.

322

Fertilization decreased the AOA to AOB ratio in the relatively fertile soils under plant canopies

323

(Fig. 1; Suppl. Table 2; N addition, 3.6 AOA/AOB; Control, 4.9 AOA/AOB) but generally

324

increased it in the interplant soils (N addition, 6.2 AOA/AOB; Control, 4.3 AOA/AOB).

325

15

326

Slurry maximum AO rates (i.e., at Vmax) were significantly higher after long-term N addition

327

compared to those of the control soils (Fig. 2; P < 0.05 in both cases). This trend was also

328

supported by measurements of static maximum AO rates (Suppl. Fig. 1; P < 0.05 in both cases).

329

NH4+ supplementation only enhanced AO rates in the static incubations of unfertilized soils but

330

not in fertilized soil (Suppl. Fig. 1), nor in any slurried incubations. These patterns show that

331

rates of AO under undisturbed, unamended conditions are NH4+-limited and may be influenced

332

by anthropogenic N additions. Taken together with the microbial abundance data, these results

333

suggest that at least some of the AOA and AOB are likely contributors to AO (Suppl. Fig 2) and

334

– as shown by the significant increases of AO rates as well as the abundance of archaeal and

335

bacterial amoA genes after N addition – both archaeal and bacterial ammonia-oxidizers are

336

responsive to environmental change.

337 338

To investigate the functional capacity of ammonia-oxidizing communities in bulk soils, we

339

evaluated the affinity (Km) parameter from the AO kinetics plots (Fig. 2; Suppl. Fig. 1). Km was

340

not possible to estimate formally in the shaken slurry assays since rates were always close to

341

maximum regardless of supplemental N addition, highlighting the low ammonia demand of

342

ammonia-oxidizers in desert soils. Residual NH4+ as low as 17 µM were measured in these

343

assays (Fig. 2), implying that the community Km is likely at or below this low value, which is

344

significantly lower than typical Km values for known AOB cultures (see rates compiled in

345

Martens-Habbena et al., 2009). In the intact incubations that were not continuously aerated

346

(static AO rates; Suppl. Fig. 1), the mean Km was 2.6 mM for control soils under plants and 1.2

347

mM for the control soils between plants. Effects of N addition could not be evaluated, given that

348

the long-term N addition itself prevented incubations at low enough ammonium.

16

349 350

An alternative way of looking at differential efficiency in ammonium utilization is to normalize

351

the maximum AO rate by the size of the community (Fig. 3; Suppl. Fig. 3). Here, AO was more

352

efficient (higher rates per copy of amoA gene) under plants than between plants (P < 0.05 in all

353

cases), and long-term N addition led to more efficient rates of AO compared to control soils in

354

the spaces between plants (significant for the shaken-slurry assay; Fig. 3; P < 0.001; Fig. 3;

355

Suppl. Fig. 3). These results suggest a change in community function (NH3 processed per amoA),

356

given that the ammonia-oxidizing community adapted favorably to higher nutrient soils (e.g.,

357

under plants and N addition).

358 359

Composition of the ammonia-oxidizing community

360 361

All phylotypes detected were related to either Nitrososphaera (Thaumarchaeota) or

362

Nitrosomonas and Nitrosospira (both β-Proteobacteria), at a ratio of about 45:10:1, respectively

363

(Fig. 4; Fig. 5). Even though this ratio is subject to potential primer biases, choosing primers that

364

are unlikely to miss abundant AOM groups (Junier et al. 2008) helps to combine qPCR and

365

sequencing data for approximate abundance comparisons of phylotypes across domains. The

366

most abundant phylotype, belonging to the Nitrososphaera subcluster 1.1, accounted for 60% of

367

all the amoA sequences. The community composition was minimally influenced by patch type or

368

N addition treatment, when assessed at the level of OTUs, and we could not detect any

369

significant differences in the relative abundance of the dominant members (Fig. 4). In soil under

370

plants, N addition decreased AOA phylogenetic diversity (PD) (P < 0.001) but increased the

371

within-group variance for AOB (PERMDISP analysis, P = 0.037), suggesting that N addition has

17

372

distinct effects on community relatedness of the AOA than of the AOB. However, all other α and

373

β diversity metrics revealed that the structure of AOA and AOB was not influenced by N

374

addition or patch type (P > 0.1 in all cases). Together – at least as far as one can detect based on

375

the amoA gene sequences – these data suggest that long-term N addition had a minor effect, if at

376

all, on AOA and AOB community structure.

377 378

Discussion

379 380

Source of N for ammonia-oxidizers

381 382

Many studies from various non-arid ecosystems have shown that inorganic N addition either

383

does not affect AOA or allows AOB to outcompete AOA (e.g., Jia & Conrad, 2009; Di et al.,

384

2009; Stopnisek et al., 2010; Xia et al., 2011; Levicnik-Hofferle et al., 2012; and reviewed in

385

Hatzenpichler, 2012). A few studies have shown that AOA may react favorably to NH3

386

originating from organic N sources or N mineralization (Chen et al., 2008; Schauss et al., 2009;

387

Kelly et al., 2011; Lu et al., 2012; Levicnik-Hofferle et al., 2012; Daebeler et al., 2012). Studies

388

showing an increase in AOA abundance after inorganic N additions are rare (Verhamme et al.,

389

2011; Daebeler et al., 2014; current study). The positive response by AOA may be explained by

390

NH3 availability from organic N or mineralization from organic sources (He et al., 2012).

391

However, organic N inputs are relatively low in ecosystems such as deserts and other extreme

392

environments (Schimel & Bennett, 2004; Booth et al., 2005). Although N inputs to arid lands

393

significantly increase productivity and N content of seasonal herbaceous annual plants, net

394

potential N mineralization in soil does not appear to be consistently augmented by N addition –

18

395

perhaps due to the frequency of water limitation, the patchiness of plant growth, and organic

396

matter loss pathways such as photodegradation and aeolian/hydrologic transport (Hall et al.,

397

2009; Rao et al., 2009; Hall et al., 2011). Regardless of the role of organic N, our results

398

highlight the unique, positive response of AOA to long-term inorganic N addition in the low

399

organic matter plant interspaces of desert soils.

400 401

The use of inorganic N fertilizers by AOA may be plausible in arid systems. Since heterotrophs

402

are likely the first to consume organic N upon metabolic activation after drought, the typical

403

pulses of resource availability imparted by fast drying/wetting cycles may force AOA to utilize

404

NH3 from inorganic N sources (Placella & Firestone, 2013). Additionally, alkaline and hot

405

environments may enhance NH4+ deprotonation, leading to NH3-gas diffusion throughout the soil

406

matrix (McCalley & Sparks, 2009; Geisseler et al., 2010). The same strains of AOA may be

407

capable of using either NH3 from organic N or inorganic N sources depending on environmental

408

conditions (He et al., 2012), as shown in vitro for the only pure AOA isolate from soil,

409

Nitrososphaera viennensis (e.g., urea; Tourna et al., 2011), and predicted in silico based on the

410

genome of a recent enrichment culture, Candidatus Nitrososphaera gargensis (Spang et al.,

411

2012).

412 413

Size, structure, and function of ammonia oxidizing communities in arid land soils

414 415

We hypothesized that long-term N addition selects for ammonia-oxidizers that are more

416

copiotrophic (lower substrate affinity, higher activity per amoA gene copy) than those in

417

unfertilized soils (Martens-Habbena et al., 2009; Prosser & Nicol, 2012). Indeed, N addition

19

418

elevated the AO rate per amoA–copy, but this effect was significant for only the least fertile parts

419

of the landscape (in soil between plants; Fig. 3) where the desert-adapted ammonia oxidizers

420

may be functioning differently after fluctuations in the environment. Differences in organic

421

compounds between soils under and away from vegetation may affect function of ammonia

422

oxidizers as it does in cultures (Lehtovirta-Morley et al., 2014). However, the relative abundance

423

of dominant amoA OTUs was constant across treatments, with small changes only in the minor

424

members (Fig. 4). Of course we cannot fully discount the idea that perhaps the minor OTUs

425

represent those that are ecologically relevant, while the numerically dominant groups are less

426

efficient or inactive (Lennon & Jones, 2011). This scenario has yet to be proven experimentally

427

and is unlikely to be the case here since archaeal and bacterial amoA gene abundance – largely

428

determined by the common OTUs – was positively correlated with AO rates (Suppl. Fig. 2).

429 430

Evidence of unique ammonia oxidation patterns in deserts

431 432

Arid land soils face extreme environmental conditions that may select for unique phylogeny and

433

niche separation. Terrestrial studies worldwide have revealed that the “marine” clade AOA

434

(Group I.1a) are often the main contributors to AO and responders to changes in conditions from

435

soil incubations (Hatzenpichler, 2012), despite being outnumbered by the “soil” clade (Group

436

I.1b; Verhamme et al., 2011; Isobe et al., 2012; Long et al., 2012; Zhang et al., 2012; Lu & Jia,

437

2013). Here, we show that AOA within the “soil” clade responded significantly to N addition,

438

and the abundance of this group was positively related to AO rates in desert soil. We also found

439

that Nitrosomonas sequences outnumbered Nitrosospira, a rarity pattern for soil systems.

440

Wastewater discharge in a desert environment was found to harbor Nitrosomonas-like strains

20

441

(Angel et al., 2010), but is an unlikely scenario for the rural location of our soils in a protected

442

state park. Alternatively, dominance of many Nitrosomonas spp. appears to be limited to

443

alkaline, high-salt, and sometimes high-NH4+ conditions (Webster et al., 2005; Koops et al.,

444

2006; Cantera et al., 2006; Ke & Lu, 2012). Pulsed resource availability – a characteristic of arid

445

lands – may also drive this distribution, since Nitrosomonas strains have advantages over

446

Nitrosospira such as faster growth responses after starvation (Bollmann et al., 2002).

447

Additionally, in most soils studied previously, AOA outnumber AOB to a greater extent than

448

found in our study (Leininger et al., 2006). In the occasional cases where AOB outnumber AOA,

449

typically up to 10-fold in terrestrial systems (e.g., Di et al., 2009), other arid lands also have a

450

novel distribution as AOB outnumber AOA by 100-fold in cold desert biocrusts (Marusenko et

451

al., 2013b). Overall, atypical ammonia-oxidizing communities appear to occupy desert soils.

452 453

Growth and activity characteristics derived from culture experiments can be combined with

454

environmental data to explore relationships between AOA and AOB at the physiological and

455

ecosystem scale (Stark & Firestone, 1996; Schauss et al., 2009; Prosser & Nicol, 2012). For

456

example, since maximum AO activity per cell is higher for Nitrosospira and Nitrosomonas

457

strains than for AOA (10 and 35-fold, respectively), the contribution of AOB to our AO rates

458

must be much more important than could be predicted from their abundance. With a 10:1 ratio of

459

abundance between Nitrosomonas and Nitrosospira in our soils, the weighted average maximum

460

cell activity for Nitrosomonas plus Nitrosospira should be 33-fold higher than that of AOA.

461

Based on the assumption that 1 amoA copy exists per AOA cell, and that a weighted average of

462

2.1 amoA copies are found per AOB cell (2 and 3 amoA copies per cell for Nitrosomonas and

463

Nitrosospira, respectively; Norton et al., 2002), we can estimate that the maximum AO activity

21

464

per amoA copy is 16-fold higher for AOB than AOA in our soils. Assuming equal number of

465

genomes and level of transcription/translation of amoA, the fact that AOA amoA copies

466

outnumber AOB in our soils by 4.3-fold still means that AOB contribute 3.7-fold more than

467

AOA to overall AO rates. These calculations are consistent with our data, which show that AOB

468

contribute on average 4.5 times more to AO rates than AOA (compare slopes in Suppl. Fig. 2).

469

Even though AOB are the dominant contributors at the ecosystem scale (e.g., total AO), the

470

doubling of AOA abundance in soils between plants of N fertilized plots means that the relative

471

importance of AOB and AOA to AO may change with N increases. This type of study refines

472

predictions of how environmental conditions affect the link between community dominance and

473

AO rates.

474 475

Conclusion

476 477

N addition affects arid land N cycling primarily through changes in community size, but less so

478

through changes in community composition. This study shows significant and positive effects of

479

inorganic N addition on abundance of Nitrososphaera-related AOA in soils. This pattern has

480

been rarely shown before, especially where N inputs from organic sources are low such as in

481

unique conditions of desert soils. Increased anthropogenic activity resulting in environmental N

482

enrichment may continue to alter ecosystem function through responses by both the AOA and

483

AOB. This work stresses the importance of research in arid lands in that results from mesic

484

systems may not be readily applicable, particularly given that agricultural and pastoral systems in

485

drylands occupy ~32% of the terrestrial land surface worldwide and often contain alkaline soil

486

that is routinely exposed to high temperatures (Koohafkan & Stewart, 2008). These systems may

22

487

contain AOM communities more similar to hot deserts than to arable lands from more mesic

488

environments.

489 490

Our results highlight the effects of N enrichment on AO rates and the community size of

491

ammonia oxidizers. We explored patterns resulting from long-term N enrichment, yet it remains

492

to be seen whether population dominance also shifts during short-term N changes associated

493

with pulsed moisture fluctuations that are characteristic of arid lands. Seasonal changes may

494

occur in AO communities (e.g. AOA abundance may dominate relative to AOB in the summer;

495

Sher et al., 2013), but it is currently unclear whether these changes are related to rates of

496

nitrification in arid and semi-arid soils following long-term N additions (Hall et al., 2011).

497

Future work is essential to investigate how our results compare to those of other arid lands and at

498

scales that were not tested here. Further research is also necessary to predict the AOM

499

contribution to ecologically and atmospherically important gases such as N2O or NO from

500

nitrifier denitrification and nitrification in these desert soils.

501 502

Acknowledgements

503 504

We would like to thank David Huber, Jennifer Learned, Brenda Ramirez, Julea Shaw and Natalie

505

Myers for assistance with lab work and training. We are grateful to Jean McLain, Egbert

506

Schwartz, Estelle Couradeau, and Elizabeth Cook for manuscript review. This work is supported

507

by NSF through the CAP LTER program (grant BCS-1026865). Funding was also provided by

508

the NSF Western Alliance to Expand Student Opportunities (WAESO) program and the GPSA at

509

ASU. The authors declare no conflict of interest.

510 23

510 511

References

512 513 514

Adair KL & Schwartz E (2008) Evidence that ammonia-oxidizing archaea are more abundant than ammonia-oxidizing bacteria in semiarid soils of northern Arizona, USA. Microb Ecol 56(3):420-426.

515 516 517

Adair KL & Schwartz E (2011) Stable Isotope Probing with 18O-Water to Investigate Growth and Mortality of Ammonia Oxidizing Bacteria and Archaea in Soil. In: Martin G. Klotz (ed) Methods in Enzymology. Academic Press, Oxford, pp. 155-169.

518 519 520

Angel R, Asaf L, Ronen Z & Nejidat A (2010) Nitrogen transformations and diversity of ammonia-oxidizing bacteria in a desert ephemeral stream receiving untreated wastewater. Microb Ecol 59(1): 46-58.

521 522

Arp DJ & Stein LY (2003) Metabolism of inorganic N compounds by ammonia-oxidizing bacteria. Crit Rev Biochem Mol Biol 38(6):471-495.

523 524 525

Austin A, Yahdjian L, Stark J, Belnap J, Porporato A, Norton U, Ravetta DA & Schaeffer SM (2004) Water pulses and biogeochemical cycles in arid and semiarid ecosystems. Oecologia 141(2):221-235.

526 527 528

Avrahami S & Bohannan BJM (2007) Response of Nitrosospira sp strain AF-Like ammonia oxidizers to changes in temperature, soil moisture content, and fertilizer concentration. Appl Environ Microbiol 73(4):1166-1173.

529 530

Belser LW & Mays EL (1980) Specific-Inhibition of Nitrite Oxidation by Chlorate and its use in Assessing Nitrification in Soils and Sediments. Appl Environ Microbiol 39(3):505-510.

531 532 533

Bollmann A, Bar-Gilissen MJ & Laanbroek HJ (2002) Growth at low ammonium concentrations and starvation response as potential factors involved in niche differentiation among ammoniaoxidizing bacteria. Appl Environ Microbiol 68(10):4751-4757.

534 535

Booth M, Stark J &Rastetter E (2005) Controls on nitrogen cycling in terrestrial ecosystems: A synthetic analysis of literature data. Ecol Monogr 75(2):139-157.

536 537

Cantera JJL, Jordan FL & Stein LY (2006) Effects of irrigation sources on ammonia-oxidizing bacterial communities in a managed turf-covered aridisol. Biol Fertility Soils 43(2):247-255.

538 539 540

Caporaso JG, Bittinger K, Bushman FD, DeSantis TZ, Andersen GL & Knight R (2010a) PyNAST: a flexible tool for aligning sequences to a template alignment. Bioinformatics 26(2):266-267.

541 542

Caporaso JG, Kuczynski J, Stombaugh J et al. (2010b) QIIME allows analysis of highthroughput community sequencing data. Nat Methods 7(5):335-336.

24

543 544

Chen X, Zhu Y, Xia Y, Shen J & He J (2008) Ammonia-oxidizing archaea: important players in paddy rhizosphere soil? Environ Microbiol 10(8):1978-1987.

545 546 547

Chu H, Fujii T, Morimoto S, Lin X, Yagi K, Hu J & Zhang J (2007) Community structure of ammonia-oxidizing bacteria under long-term application of mineral fertilizer and organic manure in a sandy loam soil. Appl Environ Microbiol 73(2):485-491.

548 549

Collins SL, Sinsabaugh RL, Crenshaw C, Green L, Porras-Alfaro A, Stursova M & Zeglin LH (2008) Pulse dynamics and microbial processes in aridland ecosystems. J Ecol 96(3):413-420.

550 551

Cook EM (2014) Direct and indirect ecological consequences of human activities in urban and native ecosystems. Dissertation. Arizona State University.

552 553 554

Daebeler A, Abell GC, Bodelier PL, Bodrossy L, Frampton DM, Hefting MM & Laanbroek HJ (2012) Archaeal dominated ammonia-oxidizing communities in Icelandic grassland soils are moderately affected by long-term N fertilization and geothermal heating. Front Microbiol 3:352.

555 556 557

Daebeler A, Bodelier PLE, Yan Z, Hefting MM, Jia Z & Laanbroek HJ (2014) Interactions between Thaumarchaea, Nitrospira and methanotrophs modulate autotrophic nitrification in volcanic grassland soil. ISME J doi:10.1038/ismej.2014.81

558 559 560

Davies R & Hall SJ (2010) Direct and indirect effects of urbanization on soil and plant nutrients in desert ecosystems of the Phoenix metropolitan area, Arizona (USA). Urban Ecosystems 13:295-317.

561 562 563

Di HJ, Cameron KC, Shen JP, Winefield CS, O'Callaghan M, Bowatte S & He JZ (2009) Nitrification driven by bacteria and not archaea in nitrogen-rich grassland soils. Nat Geosci 2(9):621-624.

564 565 566

Dowd SF, Sun Y, Wolcott RD, Domingo A & Carroll JA (2008) Bacterial tag-encoded FLX amplicon pyrosequencing (bTEFAP) for microbiome studies: Bacterial diversity in the ileum of newly weaned Salmonella-infected pigs. Foodborne Pathog Dis 5(4):459-472.

567 568

Edgar RC (2010) Search and clustering orders of magnitude faster than BLAST. Bioinformatics 26(19):2460-2461.

569 570 571

Francis CA, Roberts KJ, Beman JM, Santoro AE & Oakley BB (2005) Ubiquity and diversity of ammonia-oxidizing archaea in water columns and sediments of the ocean. Proc Natl Acad Sci USA 102(41):14683-14688.

572 573 574

Garcia-Pichel F, Lopez-Cortes A & Nubel U (2001) Phylogenetic and morphological diversity of cyanobacteria in soil desert crusts from the Colorado Plateau. Appl Environ Microbiol 67(4):1902-1910.

575 576

Geisseler D, Horwath WR, Joergensen RG & Ludwig B (2010) Pathways of nitrogen utilization by soil microorganisms - A review. Soil Biol Biochem 42(12):2058-2067.

25

577 578 579

Gelfand I & Yakir D (2008) Influence of nitrite accumulation in association with seasonal patterns and mineralization of soil nitrogen in a semi-arid pine forest. Soil Biol Biochem 40(2):415-424.

580 581

Hall SJ, Ahmed B, Ortiz P, Davies R, Sponseller RA & Grimm NB (2009) Urbanization Alters Soil Microbial Functioning in the Sonoran Desert. Ecosystems 12(4):654-671.

582 583 584

Hall SJ, Sponseller RA, Grimm NB, Huber D, Kaye JP, Clark C & Collins SL (2011) Ecosystem response to nutrient enrichment across an urban airshed in the Sonoran Desert. Ecol Appl 21(3):640-660.

585 586 587 588

Hart SC, JM Stark, EA Davidson & MK Firestone (1994) Nitrogen mineralisation, immobilisation, and nitrification, p. 985–1018. In RW Weaver, JS Angle, PS Bottomley (ed.), Methods of soil analysis. Part 2. Microbiological and chemical properties. Soil Science Society of America, Madison, Wis.

589 590

Hatzenpichler R (2012) Diversity, Physiology, and Niche Differentiation of Ammonia-Oxidizing Archaea. Appl Environ Microbiol 78(21):7501-7510.

591 592

He J, Hu H & Zhang L (2012) Current insights into the autotrophic thaumarchaeal ammonia oxidation in acidic soils. Soil Biol Biochem 55:146-154.

593 594 595

Isobe K, Koba K, Suwa Y, Ikutani J, Fang Y, Yoh M, Mo J, Otsuka S & Senoo K (2012) High abundance of ammonia-oxidizing archaea in acidified subtropical forest soils in southern China after long-term N deposition. FEMS Microbiol Ecol 80(1):193-203.

596 597

Jia Z & Conrad R (2009) Bacteria rather than Archaea dominate microbial ammonia oxidation in an agricultural soil. Environ Microbiol 11(7):1658-1671.

598 599

Johnson SL, Budinoff CR, Belnap J & Garcia-Pichel F (2005) Relevance of ammonium oxidation within biological soil crust communities. Environ Microbiol 7(1):1–12.

600 601 602

Jordan FL, Cantera JJL, Fenn ME & Stein LY (2005) Autotrophic ammonia-oxidizing bacteria contribute minimally to nitrification in a nitrogen-impacted forested ecosystem. Appl Environ Microbiol 71(1):197-206.

603 604 605

Junier P, Kim O, Molina V, Limburg P, Junier T, Imhoff JF & Witzel KP (2008) Comparative in silico analysis of PCR primers suited for diagnostics and cloning of ammonia monooxygenase genes from ammonia-oxidizing bacteria. FEMS Microbiol Ecol 64(1):141-152.

606 607

Ke X & Lu Y (2012) Adaptation of ammonia-oxidizing microorganisms to environment shift of paddy field soil. FEMS Microbiol Ecol 80(1):87-97.

608 609 610

Kelly JJ, Policht K, Grancharova T & Hundal LS (2011) Distinct Responses in AmmoniaOxidizing Archaea and Bacteria after Addition of Biosolids to an Agricultural Soil. Appl Environ Microbiol 77(18):6551-6558.

26

611 612

Konneke M, Bernhard AE, de la Torre JR, Walker CB, Waterbury JB & Stahl DA (2005) Isolation of an autotrophic ammonia-oxidizing marine archaeon. Nature 437(7058):543-546.

613 614

Koohafkan P & Stewart BA (2008) Water and cereals in drylands. The Food and Agriculture Organization of the United Nations and Earthscan.

615 616

Koops H, Purkhold U, Pommerening-Röser A, Timmermann G & Wagner M (2006) The lithoautotrophic ammonia-oxidizing bacteria. The Prokaryotes 5:778-811.

617 618 619

Koper TE, Stark JM, Habteselassie MY & Norton JM (2010) Nitrification exhibits Haldane kinetics in an agricultural soil treated with ammonium sulfate or dairy-waste compost. FEMS Microbiol Ecol 74(2):316-322.

620 621

Kowalchuk GA & Stephen JR (2001) Ammonia-oxidizing bacteria: A model for molecular microbial ecology. Annu Rev Microbiol 55:485-529.

622

Lehtovirta‐Morley LE, Ge C, Ross J, Yao H, Nicol GW & Prosser JI (2014) Characterisation of

623 624

terrestrial acidophilic archaeal ammonia oxidisers and their inhibition and stimulation by organic compounds. FEMS Microbiol Ecol. doi: 10.1111/1574-6941.12353

625 626 627

Leininger S, Urich T, Schloter M, Schwark L, Qi J, Nicol GW, Prosser JI, Schuster SC & Schleper C (2006) Archaea predominate among ammonia-oxidizing prokaryotes in soils. Nature 442(7104):806-809.

628 629

Lennon JT & Jones SE (2011) Microbial seed banks: the ecological and evolutionary implications of dormancy. Nature Rev Microbiol 9(2):119-130.

630 631 632

Levicnik-Hofferle S, Nicol GW, Ausec L, Mandic-Mulec I & Prosser JI (2012) Stimulation of thaumarchaeal ammonia oxidation by ammonia derived from organic nitrogen but not added inorganic nitrogen. FEMS Microbiol Ecol 80(1):114-123.

633 634 635

Long X, Chen C, Xu Z, Linder S & He J (2012) Abundance and community structure of ammonia oxidizing bacteria and archaea in a Sweden boreal forest soil under 19-year fertilization and 12-year warming. J Soils Sediments 12(7):1124-1133.

636 637

Low AP, Stark JM & Dudley LM (1997) Effects of soil osmotic potential on nitrification, ammonification, N-assimilation, and nitrous oxide production. Soil Sci 162(1):16-27.

638 639

Lozupone C & Knight R (2005) UniFrac: a new phylogenetic method for comparing microbial communities. Appl Environ Microbiol 71(12):8228-8235.

640 641

Lu L & Jia Z (2013) Urease gene-containing Archaea dominate autotrophic ammonia oxidation in two acid soils. Environ Microbiol 15(6):1795-1809.

27

642 643 644

Lu L, Han W, Zhang J, Wu Y, Wang B, Lin X, Zhu J, Cai Z & Jia Z (2012) Nitrification of archaeal ammonia oxidizers in acid soils is supported by hydrolysis of urea. ISME J 6(10):19781984.

645 646 647

Martens-Habbena W & Stahl DA (2011) Nitrogen Metabolism and Kinetics of AmmoniaOxidizing Archaea. Methods in Enzymology, Vol 46: Research on Nitrification and Related Processes, Pt B, Academic Press, Oxford 496:465-487.

648 649 650

Martens-Habbena W, Berube PM, Urakawa H, de la Torre JR & Stahl DA (2009) Ammonia oxidation kinetics determine niche separation of nitrifying Archaea and Bacteria. Nature 461(7266):976-979.

651 652

Marusenko Y, Huber DP & Hall SJ (2013a) Fungi mediate nitrous oxide production but not ammonia oxidation in aridland soils of the southwestern US. Soil Biol Biochem 63:24-36.

653 654 655

Marusenko Y, Bates ST, Anderson I, Johnson SL, Soule T & Garcia-Pichel F (2013b) Ammoniaoxidizing archaea and bacteria are structured by geography in biological soil crusts across North American arid lands. Ecol Proc 2(9).

656 657

McCalley CK & Sparks JP (2009) Abiotic Gas Formation Drives Nitrogen Loss from a Desert Ecosystem. Science 326(5954):837-840.

658 659

McCrackin ML, Harms TK, Grimm NB, Hall SJ & Kaye JP (2008) Responses of soil microorganisms to resource availability in urban, desert soils. Biogeochemistry 87(2):143-155.

660 661

Nishio T & Fujimoto T (1990) Kinetics of Nitrification of various Amounts of Ammonium Added to Soils. Soil Biol Biochem 22(1):51-55.

662 663

Norton JM, Alzerreca JJ, Suwa Y & Klotz MG (2002) Diversity of ammonia monooxygenase operon in autotrophic ammonia-oxidizing bacteria. Arch Microbiol 177(2):139-149.

664 665 666

Norton JM & Stark JM (2011) Regulation and Measurement of Nitrification in Terrestrial Systems. In: Martin G. Klotz (ed) Methods in Enzymology. Academic Press, Oxford, pp. 343368.

667 668

Offre P, Prosser JI & Nicol GW (2009) Growth of ammonia-oxidizing archaea in soil microcosms is inhibited by acetylene. FEMS Microbiol Ecol 70(1):99-108.

669 670 671

Okano Y, Hristova KR, Leutenegger CM, Jackson LE, Denison RF, Gebreyesus B, Lebauer D & Scow KM (2004) Application of real-time PCR to study effects of ammonium on population size of ammonia-oxidizing bacteria in soil. Appl Environ Microbiol 70(2):1008-1016.

672

Pester M, Rattei T, Flechl S, Groengroeft A, Richter A, Overmann J, Reinhold‐Hurek B, Loy A

673

& Wagner M (2012) amoA-based consensus phylogeny of ammonia-oxidizing archaea and deep 28

674 675

sequencing of amoA genes from soils of four different geographic regions. Environ Microbiol 14(2):525-539.

676 677

Placella SA & Firestone MK (2013) Transcriptional Response of Nitrifying Communities to Wetting of Dry Soil. Appl Environ Microbiol 79(10):3294-3302.

678

Prosser JI (1989) Autotrophic nitrification in bacteria. Adv Microb Physiol 30:125-181.

679 680

Prosser JI & Nicol GW (2012) Archaeal and bacterial ammonia-oxidisers in soil: the quest for niche specialisation and differentiation. Trends Microbiol 20(11):523-531.

681 682

Prosser JI & Nicol GW (2008) Relative contributions of archaea and bacteria to aerobic ammonia oxidation in the environment. Environ Microbiol 10(11):2931-2941.

683 684 685 686

Purkhold U, Pommerening-Roser A, Juretschko S, Schmid MC, Koops HP & Wagner M (2000) Phylogeny of all recognized species of ammonia oxidizers based on comparative 16S rRNA and amoA sequence analysis: Implications for molecular diversity surveys. Appl Environ Microbiol 66(12):5368-5382.

687 688

Rajeev L, da Rocha UN, Klitgord N et al. (2013) Dynamic cyanobacterial response to hydration and dehydration in a desert biological soil crust. ISME J 7:2178-2191.

689 690 691

Rao LE, Parker DR, Bytnerowicz A & Allen EB (2009) Nitrogen mineralization across an atmospheric nitrogen deposition gradient in Southern California deserts. J Arid Environ 73(10):920-930.

692 693

Rosentreter R, Bowker M & Belnap J (2007) A Field Guide to Biological Soil Crusts of Western U.S. Drylands: Common Lichens and Bryophytes. Bureau of Land Management.

694 695 696

Rotthauwe JH, Witzel KP & Liesack W (1997) The ammonia monooxygenase structural gene amoA as a functional marker: Molecular fine-scale analysis of natural ammonia-oxidizing populations. Appl Environ Microbiol 63(12):4704-4712.

697 698

Schade J & Hobbie S (2005) Spatial and temporal variation in islands of fertility in the Sonoran Desert. Biogeochemistry 73(3):541-553.

699 700

Schauss K, Focks A, Leininger S et al. (2009) Dynamics and functional relevance of ammoniaoxidizing archaea in two agricultural soils. Environ Microbiol 11(2):446-456.

701 702

Schimel JP & Bennett J (2004) Nitrogen mineralization: Challenges of a changing paradigm. Ecology 85(3):591-602.

703 704

Schimel J, Balser TC & Wallenstein M (2007) Microbial stress-response physiology and its implications for ecosystem function. Ecology 88(6):1386-1394.

705 706

Schleper C & Nicol GW (2010) Ammonia-Oxidising Archaea - Physiology, Ecology and Evolution. Adv Microb Physiol 57:1-41.

29

707 708 709

Shen J, Zhang L, Zhu Y, Zhang J & He J (2008) Abundance and composition of ammoniaoxidizing bacteria and ammonia-oxidizing archaea communities of an alkaline sandy loam. Environ Microbiol 10(6):1601-1611.

710

Sher Y, Zaady E & Nejidat A (2013) Spatial and temporal diversity and abundance of ammonia

711

oxidizers in semi‐arid and arid soils: indications for a differential seasonal effect on archaeal and

712

bacterial ammonia oxidizers. FEMS Microbiol Ecol 86(3):544-556.

713 714 715

Spang A, Poehlein A, Offre P et al. (2012) The genome of the ammonia-oxidizing Candidatus Nitrososphaera gargensis: insights into metabolic versatility and environmental adaptations. Environ Microbiol 14(12):3122-3145.

716 717 718

Sparks DL, Page AL, Helmke PA, Loeppert RH, Soltanpour PN, Tabatabai MA, Johnston CT & Sumner ME (1996) Methods of soil analysis. Part 3-Chemical methods. Soil Science Society of America, xxi-1390.

719 720

Stark JM & Firestone MK (1996) Kinetic characteristics of ammonium-oxidizer communities in a California oak woodland-annual grassland. Soil Biol Biochem 28(10-11):1307-1317.

721 722 723 724

Stephen J, Chang Y, Macnaughton S, Kowalchuk G, Leung K, Flemming C & White DC (1999) Effect of toxic metals on indigenous soil p-subgroup proteobacterium ammonia oxidizer community structure and protection against toxicity by inoculated metal-resistant bacteria. Appl Environ Microbiol 65(1):95-101.

725 726 727

Stopnisek N, Gubry-Rangin C, Hoefferle S, Nicol GW, Mandic-Mulec I & Prosser JI (2010) Thaumarchaeal Ammonia Oxidation in an Acidic Forest Peat Soil Is Not Influenced by Ammonium Amendment. Appl Environ Microbiol 76(22):7626-7634.

728 729 730

Tamura K, Peterson D, Peterson N, Stecher G, Nei M & Kumar S (2011) MEGA5: molecular evolutionary genetics analysis using maximum likelihood, evolutionary distance, and maximum parsimony methods. Mol Biol Evol 28(10):2731-2739.

731 732 733

Taylor AE & Bottomley PJ (2006) Nitrite production by Nitrosomonas europaea and Nitrosospira sp AV in soils at different solution concentrations of ammonium. Soil Biol Biochem 38(4):828-836.

734 735

Tourna M, Freitag TE & Prosser JI (2010) Stable Isotope Probing Analysis of Interactions between Ammonia Oxidizers. Appl Environ Microbiol 76(8):2468-2477.

736 737 738

Tourna M, Freitag TE, Nicol GW & Prosser JI (2008) Growth, activity and temperature responses of ammonia-oxidizing archaea and bacteria in soil microcosms. Environ Microbiol 10(5):1357-1364.

30

739 740

Tourna M, Stieglmeier M, Spang A et al. (2011) Nitrososphaera viennensis, an ammonia oxidizing archaeon from soil. Proc Natl Acad Sci USA 108(20):8420-8425.

741 742

Verhamme DT, Prosser JI & Nicol GW (2011) Ammonia concentration determines differential growth of ammonia-oxidising archaea and bacteria in soil microcosms. ISME J 5(6):1067-1071.

743

Warren A, Sud YC & Rozanov B (1996) The future of deserts. J Arid Environ 32(1):75-89.

744 745 746

Webster G, Embley T, Freitag T, Smith Z & Prosser J (2005) Links between ammonia oxidizer species composition, functional diversity and nitrification kinetics in grassland soils. Environ Microbiol 7(5):676-684.

747 748 749

Wessen E, Nyberg K, Jansson JK & Hallin S (2010) Responses of bacterial and archaeal ammonia oxidizers to soil organic and fertilizer amendments under long-term management. Appl Soil Ecol 45(3):193-200.

750 751

Widmer P, Brookes P & Parry L (1989) Microbial Biomass Nitrogen Measurements in Soils Containing Large Amounts of Inorganic Nitrogen. Soil Biol Biochem 21(6):865-867.

752 753 754

Wilhelm R, Abeliovich A & Nejidat A (1998) Effect of long-term ammonia starvation on the oxidation of ammonia and hydroxylamine by Nitrosomonas europaea. J Biochem 124(4):811815.

755 756

Winogradsky S (1890) Recherches sur les organismes de la Nitrification. Ann Inst Pasteur, Paris, 4:213-257, 760-771.

757 758

Xia W, Zhang C, Zeng X et al. (2011) Autotrophic growth of nitrifying community in an agricultural soil. ISME J 5(7):1226-1236.

759 760

Zhalnina K, de Quadros PD, Camargo FAO & Triplett EW (2012) Drivers of archaeal ammoniaoxidizing communities in soil. Front Microbiol 3:10.3389/fmicb.2012.00210.

761 762 763

Zhang L, Hu H, Shen J & He J (2012) Ammonia-oxidizing archaea have more important role than ammonia-oxidizing bacteria in ammonia oxidation of strongly acidic soils. ISME J 6(5):1032-1045.

764

31

764

Table 1. Soil characteristics from plots used in this study.

765 pH

WHC (%)a

SOM (%)b

Soil patch Treatmentd Mean SD Mean SD Mean SD typec Interplant

Control

Interplant N addition Under plant

Control

Under plant

N addition

NO2- (µg N·g-1)

NO3- (µg N·g-1)

Mean SD

Mean 1.40

SD 0.41

NH4+ (µg N·g-1) Mean

8.39 0.07 32.0 4.5

2.02 0.57

0.13 0.08

8.21 0.08 35.1 2.8

2.45 0.44

1.58 0.71

8.25 0.11 42.1 3.3

3.26 0.37

0.03 0.01

4.50

4.63

1.06

0.04

8.11 0.06 45.2 6.9

3.59 0.55

0.16 0.19

29.12

5.11

12.77

9.39

36.10 21.37

0.42

SD 0.17

22.47 21.31

Two-way ANOVA results, P value

Patch x Treatment Treatment Patch

0.651

0.992

0.868

0.171

0.459

0.464

0.007*

0.284

0.214

0.002*

0.002*

0.036*

0.034*

0.006*

0.003*

0.002*

0.772

0.519

766 767 768 769 770

a WHC = water holding capacity. b SOM = soil organic matter. c Soils were collected from spaces between plants or under the canopy of Larrea tridentata shrubs. d N addition plots were treated with 60 kg of N (as NH4NO3) ha-1·yr-1 during 2005-2012. Significance at α = 0.05 indicated by bold and *. SD = standard deviation. n = 3.

771 772

32

772

Figure legends

773 774

Figure 1. Quantification of amoA gene copy numbers for AOB and AOA from Sonoran Desert

775

soil in N addition and control plots. Error bars are standard errors of independent field triplicates.

776

33

776 777

Figure 2. Concentration-response kinetics of ammonia oxidation using the shaken-slurry assay

778

for net potential rates. To test the effect of long-term N addition on ammonia oxidation rates,

779

soils were supplemented with a range of NH4+ concentrations in the short-term laboratory

780

methods to measure kinetics of ammonia oxidation. NO2- accumulation is measured after sodium

781

chlorate inhibition as a proxy for ammonia oxidation. Bi-directional error bars are standard

782

deviations of independent field triplicates to show variation in supplemented NH4+ and measured

783

ammonia oxidation rates.

784

34

784 785

Figure 3. Effects of N addition on the function of the amoA gene-containing community using

786

estimates of copy-specific ammonia oxidation rates. Specific rates were calculated as maximum

787

ammonia oxidation rate (Vmax) from the shaken-slurry assay, divided by amoA gene copy

788

number per g soil. Error bars are standard errors of Vmax and amoA calculations for independent

789

field triplicates.

790

35

790 791

Figure 4. Community composition of OTUs (clustered at 97% nucleotide similarity) based on

792

bioinformatics of amoA gene pyrosequences. Diversity measures were analyzed separately for

793

AOA and AOB. Brackets include taxonomic classifications and percent of phylotype out of

794

AOA or AOB as an average across all treatment and patch replicates. The remaining 2% of AOB

795

were unclassified to the species-level. The remaining 2% of AOA are identified under

796

Nitrososphaera subclusters 2, 8, and 9. Despite biases with qPCR and pyrosequencing

797

technologies, the AOB and AOA bars are drawn at scale (about 25 and 75% of total amoA,

798

respectively). Each bar is the average of independent field and pyrosequencing triplicates

799

(excluding one replicate each in the interplant control and N addition samples for the AOA).

800

36

800 801

Figure 5. Neighbor-joining phylogenetic tree of archaeal and bacterial amoA gene sequences.

802

Sequences for this study were obtained from pyrosequencing of the amoA gene. Sequences of

803

known strains and subclusters are used as reference groups. OTUs were clustered at 97%

804

nucleotide similarity and taxonomy classified at multiple phylogenetic levels within the

805

Thaumarchaeota (formerly named Crenarchaeota) as proposed by Pester et al. (2012) and

806

within the AOB (Purkhold et al., 2000). The relative abundance of phylogenetic groups within

807

the AOA or AOB detected in this study is shown in parentheses next to the clade name. Height

808

of clades is proportional to the OTU richness. Bootstrap support is represented by full (75-99%)

809

and empty (50-75%) markers at the nodes.

810

37

Ammonia-oxidizing archaea respond positively to inorganic nitrogen addition in desert soils.

In soils, nitrogen (N) addition typically enhances ammonia oxidation (AO) rates and increases the population density of ammonia-oxidizing bacteria (AO...
4MB Sizes 0 Downloads 6 Views