67
Image based modeling of nutrient movement in and around the rhizosphere Author list: Keith R. Daly 1 , Samuel D. Keyes 1 , Shakil Masum 2 , Tiina Roose §1 Author affiliations: 1 Bioengineering Sciences Research Group, Faculty of Engineering and Environment, University of Southampton, University Road, SO17 1BJ Southampton, United Kingdom, 2 School of Energy, Environment and Agrifood, Cranfield University, Cranfield, MK43 0AL, United Kingdom, § Corresponding author: Faculty of Engineering and Environment, University of Southampton, Hampshire, SO17 1BJ. United Kingdom, email: [email protected] Summary Using rigorous mathematical techniques and image based modelling we quantify the effect of root hairs on nutrient uptake. In addition we investigate how uptake is influenced by growing root hairs. 1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19

eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

Embed Size (px)

Citation preview

Page 1: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

Image based modeling of nutrient movement in and around the rhizosphere

Author list: Keith R. Daly1, Samuel D. Keyes1, Shakil Masum2, Tiina Roose§1

Author affiliations:

1 Bioengineering Sciences Research Group, Faculty of Engineering and Environment,

University of Southampton, University Road, SO17 1BJ Southampton, United Kingdom,

2 School of Energy, Environment and Agrifood, Cranfield University, Cranfield, MK43 0AL,

United Kingdom,

§ Corresponding author: Faculty of Engineering and Environment, University of

Southampton, Hampshire, SO17 1BJ. United Kingdom, email: [email protected]

Summary

Using rigorous mathematical techniques and image based modelling we quantify the effect of root hairs on nutrient uptake. In addition we investigate how uptake is influenced by growing root hairs.

1

1

2

3

4

5

6

7

8

9

10

11

121314

Page 2: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

1. Abstract

In this paper we develop a spatially explicit model for nutrient uptake by root hairs based on

X-ray Computed Tomography images of rhizosphere soil structure. This work extends our

previous study (Keyes et al., 2013) to larger domains and, hence, is valid for longer times.

Unlike the model used in (Keyes et al., 2013), which considered only a small region of soil

about the root, we consider an effectively infinite volume of bulk soil about the rhizosphere.

We ask the question “at what distance away from root surfaces do the specific structural

features of root-hair and soil aggregate morphology not matter because average properties

start dominating the nutrient transport?” The resulting model captures bulk and rhizosphere

soil properties by considering representative volumes of soil far from the root and adjacent to

the root respectively. By increasing the size of the volumes we consider, the diffusive

impedance of the bulk soil and root uptake are seen to converge. We do this for two different

values of water content which provides two different geometries. We find that the size of

region for which the nutrient uptake properties converge to a fixed value is dependent on the

water saturation. In the fully saturated case the region of soil which we need to consider is

only of radius 1.1 mm for poorly soil-mobile species such as phosphate. However, in the

case of a partially saturated medium (relative saturation 0.3) we find that a radius of 1.4 mm

is necessary. This suggests that, in addition to the geometrical properties of the rhizosphere,

there is an additional effect on soil moisture properties which extends further from the root

and may relate to other chemical changes in the rhizosphere which are not explicitly included

in our model.

Keywords:

2

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

Page 3: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

Rhizosphere, X-ray CT, structural imaging, phosphate, plant-soil interaction

Abbreviations:

CT – Computer Tomography

Short title for page headings: Modelling nutrient uptake

3

39

40

41

42

43

44

45

46

Page 4: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

Introduction

The uptake of phosphate and other low mobility nutrients is essential for plant growth and,

hence, global food security (Barber, 1984; Nye and Tinker, 1977; Tinker and Nye, 2000).

However, the overuse of inorganic fertilizers has caused an accumulation of phosphate in

European soils which potentially causes eutrophication (Gahoonia and Nielsen, 2004).

Hence, there is a need to optimize the efficiency of phosphate uptake not only to increase

crop yields, but also to minimize the detrimental effects of excess phosphate on the

environment.

Plants are known to uptake nutrients both through their roots and root hairs (Datta et al.,

2011). Root hairs are thought to play a significant role in the uptake of poorly soil-mobile

nutrients such as phosphate, water uptake, plant stability and microbial interactions (Datta et

al., 2011; Gahoonia and Nielsen, 2004). In order to quantify the role of root hairs in the

uptake of poorly soil-mobile nutrients, such as phosphate, a detailed understanding of the

root-hair morphology and the region of soil around the root, known as the rhizosphere

(Hiltner, 1904), is required. This is because the properties of soil in the rhizosphere are

thought to be significantly different due to soil microbial interactions and compaction of soil

by the root (Aravena et al., 2010; Aravena et al., 2014; Daly et al., 2015; Dexter, 1987;

Whalley et al., 2005).

Early models for nutrient movement in soil treated the rhizosphere as a volume averaged

continuum (Barber, 1984; Ma et al., 2001; Nye and Tinker, 1977). More recently models for

nutrient movement have been derived and parametrized for dual porosity soil (Zygalakis et

4

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

66

67

68

69

Page 5: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

al., 2011) and for soil adjacent to cluster roots (Zygalakis and Roose, 2012). These models

were based around the technique of homogenization (Hornung, 1997; Pavliotis and Stuart,

2008), a multi-scale technique which enables field scale equations to be derived and

parametrized based on underlying continuum models.

To obtain a better understanding of processes occurring in the rhizosphere, in comparison to

the surrounding bulk soil, non-invasive measurements of the plant root and soil structure are

essential (George et al., 2014; Hallett et al., 2013). Three dimensional imaging of plant roots

in-situ using X-ray Computed Tomography (CT) is a rapidly growing field (Aravena et al.,

2010; Aravena et al., 2014; Hallett et al., 2013; Keyes et al., 2013; Tracy et al., 2010). Using

these techniques sub-micron resolutions can be achieved enabling the root hair morphology

to be visualized (Keyes et al., 2013).

In addition to direct in-situ visualization of plant-soil interaction, the use of X-ray CT also

provides the means to apply numerical models describing the diffusion of nutrients directly to

the imaged geometries (Aravena et al., 2010; Aravena et al., 2014; Daly et al., 2015; Keyes et

al., 2013; Tracy et al., 2015). This approach has been used to quantify the effectiveness of

root hairs in saturated soil conditions (Keyes et al., 2013) using a diffusion model originally

developed for mycorrihzal fungi (Schnepf et al., 2011). However, the study in (Keyes et al.,

2013) considered a small volume of soil adjacent to the root. This volume extended c.

300 μm from the root surface and had a zero flux boundary condition on the outer surface.

Hence, once the phosphate immediately adjacent to the root was depleted the uptake of

nutrients into the root stopped. This, whilst accurate for the experimental situation

considered, limited the time frame for which the model was applicable to c. 3 hours.

5

70

71

72

73

74

75

76

77

78

79

80

81

82

83

84

85

86

87

88

89

90

91

92

93

Page 6: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

In this paper we extend the work of (Keyes et al., 2013) to include both bulk and rhizosphere

soil. The key aims of this study are to develop an image based modelling approach which is

applicable to root hairs surrounded by a large/infinite bulk of soil, to understand how root

hairs contribute to nutrient uptake at different soil water content and to estimate the effects of

root hair growth on nutrient uptake. We have chosen to model a single root in a large/infinite

volume of soil, rather than multiple roots competing for nutrients for a number of reasons.

Firstly, we want to compare our model to established models for nutrient uptake which are

applied in this geometry (Roose et al., 2001). Secondly, whilst some models consider root-

root competition, (Barber, 1984), these models are applied in a cylindrical geometry. Hence,

any competition is provided by a ring of roots at a distance from the root under consideration.

In order to accurately capture roots at a given root density additional assumptions and

approximations would be required. Finally, our aim is to study the role of root hairs on the

uptake of a single root. Hence, we have chosen to study a geometry which captures and

isolates this effect. The modelling strategy developed in this paper will provide a framework

for future modelling studies to consider how different root hair morphologies influence

uptake. In addition it will provide a new level of understanding in to the role or root hairs in

nutrient uptake as the root system begins to develop.

The bulk soil properties are derived using multi-scale homogenization combined with image

based modelling in a similar manner to (Daly et al., 2015). The resulting equations are

solved analytically and are patched to an explicit image based geometry for the rhizosphere

using a time dependent boundary condition. Reducing the bulk soil to a single boundary

condition allows us to capture the full geometry and topology of the plant-root-soil system

6

94

95

96

97

98

99

100

101

102

103

104

105

106

107

108

109

110

111

112

113

114

115

116

117

Page 7: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

without the need for excessive computational resources and has particular relevance to poorly

soil-mobile species such as phosphate, potassium and zinc. The model describes phosphate

uptake by roots and root hairs in an effectively infinite volume of bulk soil whilst capturing at

high precision any changes in the soil adjacent to the plant roots. In this case an effectively

infinite volume of soil corresponds to any volume of soil which is sufficiently large that the

phosphate depletion region about the root does not reach the edge of the domain considered.

Using this model we are able to parameterize upscaled models for nutrient motion in soil

(Zygalakis et al., 2011).

This paper is arranged as follows: in section 2 we describe the plant growth, imaging and

modelling approaches, in section 3 we discuss the results of numerical simulation and show

how these models can be used to perform an image based study of root hair growth, finally in

section 4 we discuss our results and show how these models might be further developed. The

technical description of the mathematical models used in this paper is provided in appendix A

and appendix B.

2. Methods

2.1. Plant growth

We study a rice genotype Oryza Sativa cv. Varyla provided by the Japan International

Research Centre for Agricultural Sciences (JIRCAS). The seeds were heat-treated at 50℃ for 48h to break dormancy and standardise germination. Germination was carried out at

23±1℃ for 2 days between moistened sheets of Millipore filter paper. The growth medium

7

118

119

120

121

122

123

124

125

126

127

128

129

130

131

132

133

134

135

136

137

138

139

140

Page 8: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

was a sand-textured Eutric Cambisol soil collected from a surface plot at Abergwyngregyn,

North Wales (53o14’N, 4o01’W), the soil organic matter was 7%, full details are given in

(Lucas and Jones, 2006), soil B. The soil was sieved to <5 mm, autoclaved and air dried at

23±1℃ for 2 days. The dried soil was sieved to between bounds of 1680 μm and 1000 μm,

producing a well-aggregated, textured growth medium.

Growth took place in a controlled growth environment (Fitotron SGR, Weiss-Gallenkamp,

Loughborough, UK) for a period of 14 days. Growth conditions were 23±1℃ and 60 %

humidity for 16 hours (day) and 18±1℃ and 55 % humidity for 8 hours (night) with both

ramped over 30 minutes.

The seminal roots were guided by a specially designed growth environment fabricated in

ABS plastic using an UP! 3D printer (PP3DP, China). The resulting assay is shown in

Supplementary Figure 1 and further details are provided in (Keyes et al., 2015) for details.

The morphology of the growth environment was used to guide the roots into 7 syringe barrels

(root chambers) which can then be detached for imaging once the growth stage is complete.

The lower portion including the root chambers was housed in a foil-wrapped 50 ml centrifuge

tube which occludes light during the growth period.

2.2. Imaging

Imaging of root chambers was conducted at the TOMCAT beamline on the X02DA port of

the Swiss Light Source, a 3rd generation synchrotron at the Paul Scherrer Institute, Villigen,

Switzerland. The imaging resolution was 1.2 μm, and a monochromatic beam was employed

with an energy of 19 kV. A 90 ms exposure time was employed to collect 1601 projections

8

141

142

143

144

145

146

147

148

149

150

151

152

153

154

155

156

157

158

159

160

161

162

163

Page 9: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

over 180 degrees with a total scan time of 2.4 minutes. The data were reconstructed to 16-bit

volumes using a custom back-projection algorithm implementing a Parzen filter. The

resulting volume size was 2560x2560x2180 voxels.

The reconstructed volumes were analysed using a multi-pass approach (Keyes et al., 2015).

Whilst root hairs can be visualised in situ for dry soils they are challenging to distinguish

from some background phases even at high resolution. Specifically, the portions of root hairs

which traverse water filled regions of the pore space are indistinguishable from the water.

The visible sections of the root hairs were extracted manually using a graphical input tablet

(Cintiq 24, WACOM) and a software package that allowed interpolation between lofted

cross-sections (Avizo FIRE 8, FEI Company, Oregon, USA). A number of these sections

represent an incomplete root hair segmentation since proportions of the hair paths can be

occluded by fluid rich phases and pore water. Full root hair paths were extrapolated from the

segmented sections of partially visible hairs using an algorithm that simulates root hair

growth (Keyes et al., 2015). The algorithm was parameterised from the root hairs of plants

(Oryza Sativa cv. Varyla) which were grown in rhizo-boxes with dimensions 30x30x2 cm,

constructed of transparent polycarbonate with a removable front-plate, and wrapped in foil to

occlude light. Hydration during the growth period was maintained via capillary rise from a

tray of water, with occasional top-watering. The plants were grown in a glasshouse, with

temperature ranging from 20-32 ℃ over the growth period. The root systems of 1 month old

specimens were gently freed from soil using running water. One fine primary and one coarse

primary were randomly sampled before manual removal of the remaining soil from the roots.

Each was cut into a number of subsections of 2 cm length for analysis, each being imaged

using an Olympus BX50 optical microscope. After stitching all images together 5 sub-regions

were randomly defined for each section. Each region was chosen to represent a distance of 1

9

164

165

166

167

168

169

170

171

172

173

174

175

176

177

178

179

180

181

182

183

184

185

186

187

188

Page 10: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

mm on the root surface. All hairs in sub-regions which were in sharp focus were measured in

FIJI using a poly-line tool to generate a set of 220 hair lengths. A normal distribution was

fitted to these data and used to parameterize the hair-growth algorithm. The fitted and

sampled distributions are shown in Supplementary Figure 1. The mean of experimental

lengths was 171.56 µm, compared to the mean fitted lengths: 173.67 µm. The standard

deviation of the experimental lengths was 72.88 µm compared to the standard deviation of

the fitted lengths 73.44 µm.

We observe that root hair density measured via Synchrotron Radiation X-ray CT data is

generally slightly higher than for microscopy, probably due to the lack of damage artefacts.

For the data in this study, the density is ≈ 77 hairs mm−1, compared to 64 hairs mm−1 measured

via bright-field microscopy in a study of similar rice varieties grown under similar conditions

(Wissuwa and Ae, 2001). This discrepancy is, in part, because the in situ hairs are often part-

occluded by fluid and colloidal phases, it is usually only possible to obtain partial

measurements in the Synchrotron CT data. This is offset to some degree by the lack of

damage artefacts when compared to root-washing and microscopy, but the length values are

low compared to most literature sources. However, in this study we explore the implications

of soil domain size on nutrient uptake and have used hair data attained via quantification of

soil-grown roots using bright field microscopy.

The root geometry was extracted in the same manner, manually defining the root cross-

section from slices sampled (at a spacing of 120 µm) along the root growth axis. The soil

minerals, pore gas and pore fluid were extracted using a trainable segmentation plugin in FIJI

10

189

190

191

192

193

194

195

196

197

198

199

200

201

202

203

204

205

206

207

208

209

210

211

Page 11: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

that implements the WEKA classifiers for feature detection using a range of local statistics as

training parameters (Hall et al., 2009).

The segmented geometry was used to produce 3D surface meshes, using the ScanIP software

package (Simpleware Ltd., Exeter, United Kingdom). The root, hairs, gas fluid phase and

soil minerals were exported as separate .STL files which provide the geometries on which the

numerical models can be tested.

2.3. Modelling

We model the uptake of phosphate by a single root and root hairs in an unbounded partially

saturated geometry, see Figure 1. In order to do this we extend the model described in

(Keyes et al., 2013). In (Keyes et al., 2013) the authors considered a segment of root and

surrounding soil. On the outer boundary of the segment a zero flux condition was applied.

By definition this condition prevents any influx of nutrient from outside the domain

considered. Hence, the total possible uptake is limited to the nutrient available from within

the initial segment. This was adequate for the experimental situation considered, i.e., a pot of

radius c. 300 μm, but is clearly not applicable for isolated roots in soil columns of radius

larger than 300 μm . We note that this is a slightly different scenario from the one studied by

Barber (Barber, 1984) in which the domain around the root was considered to be half the size

of the inter-root spacing. In other words, Barbers model considered multiple roots effectively

introducing root competition. In this work we consider a single isolated root and how that is

affected by soil structural properties.

11

212

213

214

215

216

217

218

219

220

221

222

223

224

225

226

227

228

229

230

231

232

233

234

Page 12: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

In order to overcome this limitation we consider a single root in a large/infinite domain of

soil. However, as considering an infinite domain explicitly is computationally unfeasible, we

break the domain into two regions: the rhizosphere and bulk soil. We assume that the bulk

soil, which is far from the root, may be considered as a homogeneous medium with an

effective phosphate diffusion constant. The effective diffusion properties of phosphate in the

bulk soil are derived from the segmented CT data using the method of homogenization

(Pavliotis and Stuart, 2008; Zygalakis et al., 2011; Zygalakis and Roose, 2012). The nutrient

flux in the bulk soil is then patched to the rhizosphere domain using a boundary condition

which simulates an effective infinite region of bulk soil beyond the rhizosphere. In this

section we introduce the equations and summarize the model which results from these

assumptions. Full details of the derivation have been included in appendices A and B.

We assume that phosphate can only diffuse in the fluid domain, i.e., no diffusion occurs in

the air filled portion of the pore space. Therefore, we only present equations in the fluid

region and its associated boundaries. However, we make no assumptions about the

geometrical details regarding the air filled portion of pore space. If, from the CT data, there

is air in contact with soil or root material this simply acts to reduce the surface area over

which phosphate can be exchanged. We define two regions of fluid filled pore space we

denote the rhizosphere by Ωr as the region explicitly considered near to the root. Physically

we may think of this as an approximation to the rhizosphere. The second region is Ωb, the

region of bulk soil outside of the rhizosphere, see Figure 2. As different boundary conditions

need to be applied on each of the pore water interfaces we adopt the following naming

convention. We take the interface between the two regions, located at a distance ~rb from the

centre of the root to be Γrb. The root and root-hair surfaces are defined as Γ0 r, and Γ0 h

12

235

236

237

238

239

240

241

242

243

244

245

246

247

248

249

250

251

252

253

254

255

256

257

258

Page 13: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

respectively, the soil particle surface to be Γ sr and Γ sb in region Ωr and Ωb respectively.

Finally we define the air water interfaces as Γ ar and Γ ab in region Ωr and Ωb respectively.

The method we use to determine phosphate movement is different in each region. Phosphate

movement in the rhizosphere is calculated based on a spatially explicit model obtained from

CT data. Phosphate movement in the bulk soil is calculated analytically based on the solution

to a spatially explicit numerical model in a representative volume of bulk soil. We present

each of these models separately.

2.3.1. Rhizosphere

In the rhizosphere we assume that phosphate moves by diffusion only

∂~C r

∂ ~t=~D~∇2~C r ,

x∈Ωr , (1)

where ~Cr is the phosphate concentration in the soil solution and ~D is the diffusion constant of

phosphate in the soil solution. The phosphate is assumed to bind to the soil particles based on

linear first order kinetics

~D n⋅~∇~C r=−~ka~Cr+

~kd~Ca , x∈ Γsr , (2)

∂~C a

∂ ~t=~ka

~C r−~kd

~Ca ,x∈ Γsr , (3)

where ~k a and ~k d are the adsorption and desorption rates respectively, ~Ca is the nutrient

concentration on the soil surface and n is the outward pointing surface normal vector. Here

we have assumed that the nutrient concentration of the soil can be represented as a surface

concentration and that any replenishment from within the soil aggregate is either so fast that

13

259

260

261

262

263

264

265

266

267

268

269

270

271

272

273

274

275

276

Page 14: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

it is captured by these equations or so slow that we can neglect it. We assume that there is no

nutrient diffusion across the air water boundary

~D n⋅~∇~C r=0 , x∈ Γar . (4)

We consider uptake of phosphate from the fluid only. On the root surface this follows a

linear uptake condition

~D n⋅~∇~C r=~λ~C r , x∈ Γ0 r . (5)

where ~λ is the nutrient uptake rate on the root and root-hair surfaces. For comparison with

the paper by Keyes et al (Keyes et al., 2013) we have chosen the absorbing power as ~λ=~Fm~Km

,

consistent with Nye and Tinker. Equation (5) is the small concentration equivalent of the

Michaelis Menten condition, in which ~Fm is the maximum rate of uptake and ~Km is the

concentration when the uptake is half the maximum.

The flux on the root hair surface is given by one of two different scenarios: the first is linear

uptake

~D n⋅~∇~C r=~λ~C r , x∈ Γ0 h . (6)

The second case we consider is a pseudo time-dependent uptake which simulates linear

uptake for a growing root-hair. We have termed this pseudo root-hair growth as we do not

consider the geometrical and mechanical effects of root hair growth explicitly. Rather, we

14

277

278

279

280

281

282

283

284

285

286

287

288

289

290

291

292

293

294

Page 15: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

assume that at ~t=0 the root hairs are present, but do not contribute to uptake. As the

simulation progresses the root hairs are assumed to grow at a certain rate. We model this by

assuming an active region of root hair which grows outward from the root; this is illustrated

in Figure 3. We write the root-hair uptake condition as

~D n⋅~∇~C r=~λ2~C

r{1+ tanh [−~α (~r−~gr

~t ) ]}, x∈ Γ0h , (7)

where ~r is the axial distance along the root hair, ~gr is the root hair growth rate, ~α is an

empirical parameter which controls the sharpness of the transition between the root hairs

being “on”, i.e., actively taking up nutrient and “off”, i.e., not actively taking up nutrient.

The adsorption and desorption constants (~k a∧~k d) are taken from (Keyes et al., 2013) where

they were derived using standard soil tests (Giesler and Lundström, 1993; Murphy and Riley,

1962). The remaining parameters used in equations (1) to (5) were obtained from the

literature (Barber, 1984; Tinker and Nye, 2000) and are summarized in Table 1.

2.3.2. Rhizosphere bulk interface

Due to the success of averaged equations in bulk soil, for example Darcy’s law and Richards’

equation (Van Genuchten, 1980), the precise details of the geometry are less important in

determining the total phosphate movement and average properties are sufficient to be

considered. Therefore, rather than consider the explicit details of the geometry we derive an

effective diffusion constant based on a representative soil volume in absence of roots. This is

achieved using the method of homogenization, (Pavliotis and Stuart, 2008), which is

15

295

296

297

298

299

300

301

302

303

304

305

306

307

308

309

310

311

312

313

314

315

Page 16: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

implemented as follows. First it is shown that, on the length scale of interest, the nutrient

concentration is only weakly dependent on the precise structure of the soil. Secondly a set of

equations, often called the cell problem, are derived which determine the local variation in

concentration due to the representative soil volume. Finally an averaged equation is derived

which describes the effective rate of diffusion in terms of an effective diffusion constant

~Deff =~D D eff where Deff is calculated from the bulk soil geometry (equations (A31) to (A34)

and (A58)) and describes the impedance to diffusion offered by the soil. Full details of how

Deff is derived are provided in Appendix A. Hence, in the bulk the diffusion of phosphate is

described by

∂~Cb

∂ ~t=~Deff

~∇2~Cb ,(8)

where ~Cb is the nutrient concentration in the bulk soil. The advantage of equation (8) is that

we do not need to explicitly consider the soil geometry in the bulk soil domain. In itself this

significantly reduces the computational cost for finite but large domains. However, rather

than solving equation (8) numerically we find an approximate analytic solution for an infinite

bulk soil domain subject to the conditions ~Cb=~Cr ( t ) for x∈ Γrb and ~Cb →~C∞ as ~r →∞. Using

this solution we are able to define a relationship between concentration and flux at the edge

of the rhizosphere. The result is a condition which simulates the presence of an infinite

region of bulk soil at the rhizosphere soil domain boundary

~D n⋅~∇~C r=−2~Deff (

~C∞−~C r)

~rb ln ( 4~D eff~t

~rb2 e−γ )

,x∈ Γrb , (9)

16

316

317

318

319

320

321

322

323

324

325

326

327

328

329

330

331

332

333

334

Page 17: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

where γ=0.57721 is the Euler–Mascheroni constant. Hence, the bulk soil is dealt with

entirely by condition (9). This significantly reduces the computational cost whilst allowing

us to include the averaged geometric details of the bulk soil through the parameter ~Deff . We

note that the boundary condition (9) is singular at ~t=0. This is regularized by the fact that

when ~t=0 we have ~C∞=~C r. To overcome the difficulties of implementing this we follow the

suggestion of (Roose et al., 2001) and modify equation (9) such that for small ~t the equation

is non-singular, see Appendix B. To summarise the final set of equations we solve in the

rhizosphere are

∂~C r

∂ ~t=~D~∇2~C r ,

x∈Ωr , (10)

~D n⋅~∇~C r=−~ka~Cr+

~kd~Ca , x∈ Γsr , (11)

∂~Ca

∂ ~t=~ka

~C r−~kd

~Ca ,x∈ Γsr , (12)

~D n⋅~∇~C r=0 , x∈ Γar , (13)

~D n⋅~∇~C r=~λ~C r , x∈ Γ0 r , (14)

~D n⋅~∇~C r=−2~Deff (

~C∞−~C r)

2 ~rb+~rb [γ−ln (4~Deff~t /~r b

2+eγ ) ]x∈ Γrb, (15)

where, in the linear uptake case, the root hair boundary condition is given by

~D n⋅~∇~C r=~λ~C r , x∈ Γ0 h . (16)

In the root hair growth scenario the root hair boundary condition is given by

~D n⋅~∇~C r=~λ2~C

r{1+ tanh [−~α (~r−~gr

~t ) ]}, x∈ Γ0 h , (17)

and ~D n⋅~∇~C r=0 is applied on the remaining external boundaries, see Figure 2. Equations

(10) to (17) describe the uptake of phosphate by roots and root hairs from an infinite bulk of

soil.

17

335

336

337

338

339

340

341

342

343

344

345

346

347

348

Page 18: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

Using the STL files generated from the CT data a computational mesh is constructed using

the snappyHexMesh package. The mesh is generated from a coarse hexahedral mesh and the

STL surface meshes from the imaged geometry. The hexahedral mesh is shaped like a

segment of initial angle ~θ, height ~h and radius ~rb corresponding to the size of the domain to

be modelled. Each of these parameters is increased until the root uptake properties converge.

Successive mesh refinements are made where the mesh intersects any geometrical feature

described by the STL files, i.e., the mesh is refined about the surfaces of interest. Once the

mesh has been refined to hexahedra of side length less than 1 μm about the STL surfaces the

regions outside of the water domain are removed. The remaining mesh is then deformed to

match to the STL surfaces. Finally the mesh is smoothed to produce a high quality mesh on

which the numerical models can be run. The equations are discretised on the mesh and

solved in OpenFOAM, an open source finite volume code (Jasak et al., 2013), using a

modified version of the inbuilt LaplacianFOAM solver to couple the bulk concentration to

the sorbed concentration at the soil boundaries. Time stepping is achieved using an implicit

Euler method with a variable time stepping algorithm for speed and accuracy. All numerical

solutions were obtained using the Iridis 4 supercomputing cluster at the University of

Southampton.

3. Results and discussion

Using the theory developed in section 2.3 we calculate the bulk soil effective diffusion

constant and nutrient uptake properties of a single root with hairs. We consider two different

water contents. The first of which is the case in which all the pore space is full of water, i.e.,

S=1 where S is the volumetric water content defined as the volume of water divided by the

18

349

350

351

352

353

354

355

356

357

358

359

360

361

362

363

364

365

366

367

368

369

370

371

372

Page 19: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

volume of pore space. Secondly we use the segmented CT image to obtain the water content

and air-water interface from the scanned soil. In this case the volumetric water content is

S=0.33. In addition to being able to parametrize existing models we also calculate the size

of the region which needs to be considered for the uptake predicted by our simulations to

converge. Finally we consider how a growing root-hair system affects the overall uptake

properties of the root and root-hairs.

3.1. Bulk soil properties

Before we consider the nutrient uptake properties of the plant root system we first find the

homogenized properties of the bulk soil. From the CT image, Figure 1, we select a cube of

soil of side length Lmax=2 mm. From this we sub sample a series of geometries of size

L=2−n/3 Lmax for n=0,1 , …8, i.e., we repeatedly half the volume of bulk soil considered, and

solve equations (A31) to (A34) and (A58) to obtain ~Deff . The key difficulty that arises in

solving these equations is that they require the geometry to be periodic, i.e, it is composed of

regularly repeating units. In reality this is not the case. Hence, we have to impose periodicity.

Following the method used in (Tracy et al., 2015) we impose periodicity by reflecting the

geometry about the three coordinate axes. We note that this reflection is achieved

mathematically through analysis of the symmetries of the problem rather than by physically

copying the meshes. This is discussed in more detail in the appendix and results in solving

equations (A35) to (A39). Computational resources ranged from 46 Gb across 2 nodes for 3

minutes for Saturated case with n=8 to 900Gb across 16 nodes for 45 minutes for the more

complex partially saturated case with n=0. As we increase the size of L we find that the

value of ~Deff converges to a fixed value once a sufficiently large sub sample volume is

included. Figure 4 shows these values for saturated and partially saturated soil. We see that

19

373

374

375

376

377

378

379

380

381

382

383

384

385

386

387

388

389

390

391

392

393

394

395

396

Page 20: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

for small L the effective diffusion coefficient is a function of L. However, as L is increased

the value is seen to level off and become effectively independent of our choice of L. We

interpret this to mean that ~Deff has converged and will not change if L is increased further.

As a measure of convergence we assume that ~Deff has converged if doubling the volume of

soil considered, corresponding to increasing n by 1, causes a change in ~Deff which is smaller

than 5% of its final value. The effective diffusion constant returns a converged value for n≤ 5,

corresponding to L=0.63 mm, see Figure 4. However, due to the symmetry reduction used in

deriving equations (A35) to (A39) this side length must be doubled to obtain the true

representative volume. Hence, the actual size of the representative volume is L=1.26 mm.

At this value the computational resources used were 75Gb across 2 nodes for 45 minutes in

the saturated case and 150 Gb across 4 nodes for 7 minutes in the partially saturated case.

The larger resource requirements are necessary for the partially saturated case due to the

increased resolution needed to capture thin water films. We note that, although a

representative side length of L=1.26 mm may seem small the soils used here were sieved to

between two relatively close tolerances 1.00 and 1.68 mm. Hence, these soils will be very

homogeneous and we would expect the final soil packing to be close to an ideal sphere

packing. The value of ~Deff is seen to converge to ~Deff ≈ 0.56 ×10−9m2 s−1 for the saturated soil

and ~Deff ≈ 0.15 ×10−9m2 s−1 for the partially saturated soil. No data is plotted for the partially

saturated case with n=8 because, for samples this small, the air films and soil provided a

complete barrier to diffusion and no effective diffusion coefficient could be calculated.

3.2. Nutrient uptake properties

Using the values obtained for ~Deff at full and partial saturation we now consider the uptake of

nutrient by the plant root system shown in Figure 4. We are interested in finding the size of

20

397

398

399

400

401

402

403

404

405

406

407

408

409

410

411

412

413

414

415

416

417

418

419

420

Page 21: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

the volume of soil about the root which we need to consider geometry explicitly in order to

accurately represent the uptake of phosphate by the root and root hairs. To determine the size

of this volume we consider a segment of angle ~θ, height ~h and radius ~rb centred on the root.

The maximum radius considered is ~rmax=1.76 mm, ~θmax=π and ~hmax=1.8 mm. The height,

radius and angle of the segment considered are increased from 20% of the maximum domain

size in steps of 20% till the maximum is reached corresponding to a total of 125 simulations

for each of the saturated and partially saturated cases. Computational requirements ranged

from 15 Gb across 2 nodes for 15 minutes in the simplest case to 300 Gb across 12 nodes for

16 hours in the most complex case.

Typical plots for the nutrient movement in the saturated and partially saturated cases are

shown in Figure 5 for a variety of ~rb values with the smallest ~θ and ~h values used. Initially

we expected that there would be two convergence criteria; firstly a sufficiently large root

surface area will need to be considered in order for the number of root hairs involved in

nutrient uptake to be representative, i.e., the density of root hairs in the sub volume is equal to

the density of root measured density of root hairs. Secondly the radius of the region

considered must be sufficiently large that all the rhizosphere soil is captured. However,

whilst there is some variance in the data with root surface area, Figure 7 and error bars in

Figure 6, it seems that the main convergence criteria is that a sufficiently large value of ~rb,

i.e., the location of the outer rhizosphere-bulk soil boundary, is used. Whilst this conclusion

is somewhat surprising it seems that only a relatively small value for ~θ and ~h are required to

capture enough of the root hair structure that it is representative. On inspection we see that

the ratio between the root hair area and root area quickly settles to ≈ 0.5 ± 0.05 for root hair

surface areas greater than 0.2 mm2. Hence, for root surfaces areas above this value the

21

421

422

423

424

425

426

427

428

429

430

431

432

433

434

435

436

437

438

439

440

441

442

443

444

Page 22: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

effective density of root hair surface area does not change and the uptake properties are not

expected to change significantly.

To check the convergence of the simulated phosphorous dynamics we compare the total flux

at the root and root hair surfaces as a function of time for the linear uptake conditions. These

data are presented in Figure 6 for both saturated and partially saturated conditions for the root

uptake and Figure 7 for the root-hair uptake. The simulated nutrient uptake is seen to settle to

a rate which is independent of ~rb for sufficiently large ~rb. We assume that the total uptake

has converged once increasing ~rb produces a change in uptake of less than 5% of the final

value. The radius at which this is seen to occur is ~rb=0.6 ~rmax=1.1 mm for the saturated case

and ~rb=0.8 ~rmax=1.4mm for the partially saturated case. In the saturated case at ~rb=0.2~rmax

the cumulative uptake of the root and root hair surfaces is 58.8 ×10−3 μmol mm−2, this

converges to 50.0 ×10−3 μmol mm−2± 10−3 μmol mm−2 for ~rb ≥ 0.6 ~rmax. In the partially saturated

case the cumulative uptake at ~rb=0.2~rmax is 62.4 × 10−3 μmol mm−2 . This converges to

31 ×10−3 μmolmm−2 ±10−3 μ mol mm−2 for ~rb ≥ 0.8 ~rmax. It is interesting at this point to

compare our model with the zero flux approximation used in (Keyes et al., 2013). In this

case the total uptake can easily be calculated as the total amount of phosphate available in the

geometry at time ~t=0. For the largest saturated domain we have considered this would be a

total uptake of 5.54 ×10−4 μmolmm−2, which is significantly less than the uptake measured

here (50.0 ×10−3 μmolmm−2). We note that the uptake by the root hairs converges in a

diminishing oscillatory manner, i.e., the uptake first increase with ~rb before decaying to a

steady state. This is because at the smallest value of ~rb the geometry does not contain the

entire length of the root hairs. Hence, the root hair uptake is noticeably lower than the cases

22

445

446

447

448

449

450

451

452

453

454

455

456

457

458

459

460

461

462

463

464

465

466

467

Page 23: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

~rb ≥ 0.4 ~rmax. Once the full hair length is taken into account the convergence is smooth and

behaves the same way as the convergence of the root uptake.

As with the effective diffusion constant we find that we do not need to consider a large

amount of soil in order to obtain converged properties. Hence, we can observe that the radius

required to capture the behaviour of the rhizosphere was dependent on the saturation of the

fluid. In the fully saturated case the radius simply needed to be large enough to slightly

greater than 100% of the root hair length, i.e., 1.1 mm. However, in the partially saturated

case a much larger radius (roughly double the root hair length) was required to capture the

corresponding effect on the nutrient motion. We observe that if we using this model we can

parameterise mathematically simpler models such as the one developed in (Roose et al.,

2001). By using the effective diffusion constant and additional uptake provided by the root

hairs to parameterize the model in (Roose et al., 2001) we obtain a cumulative uptake of

50.0 ×10−3 μmolmm−2 using  ~Deff =0.56~D and ~λeff=1.3 ~λ for the uptake parameter in the

saturated case. In the partially saturated case we obtain a cumulative uptake of 31.2

×10−3 μmolmm−2 using  ~Deff =0.15~D and ~λeff=0.75~λ. In both cases the model developed by

Roose can be agrees well with our predictions assuming the effective uptake parameter is a

function of saturation. In the fully saturated case the root hairs offer an increase in uptake of

30%. In the partially saturated case the effective uptake is decreased. This is in part due to

the decrease in root and root hair surface area which is in contact with water, 80% compared

to the fully saturated case. Hence, whilst the root hairs considered in this geometry do

increase root uptake when comparing image based simulations, they do not significantly alter

the uptake parameters in the conventional models.

23

468

469

470

471

472

473

474

475

476

477

478

479

480

481

482

483

484

485

486

487

488

489

490

491

Page 24: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

3.3. Root hair growth

We now consider the root hair growth scenario. We neglect the geometrical growth of the

root hairs and consider a growing region on which uptake occurs. We use a representative set

of values for ~θ, ~h and ~rb for which the simulation has converged. Specifically we take

~θ=0.6 ~θmax, ~h=0.6 ~hmax and ~rb=0.6 ~rmax for the fully saturated case and ~θ=0.6 ~θmax,

~h=0.6 ~hmax

and ~rb=0.8 ~rmax for the partially saturated case. The root hair growth is shown schematically

in Figure 8 for a range of different times. The hair growth parameter ~gr=5 ×10−9 ms−1 is

chosen to give the root hairs an effective growth period of two days and the empirical

parameter ~α=0.2× 106 m−1 is chosen such that the ~α−1, i.e., the distance between the active

and non-active sections of the root hair is of length 5 μm. Alongside the uptake results for the

saturated and partially saturated cases, shown in Figure 9, we have plotted the uptake results

for the standard fixed root hairs case. Interestingly, other than the magnitude of the flux,

there is little qualitative difference between the saturated and partially saturated cases.

The main difference between the root hair growth and the static root hair scenario is observed

towards the start of the growth period, between 0 and 48 hours. In the growth scenario, as

expected, the uptake is dominated by the root for short times. However, as time progresses

the root-hairs grow and provide the dominant contribution to nutrient uptake. Once the

nutrients in the region immediately adjacent to the root have been taken up the uptake in the

growing root hair quickly settles to match the uptake rate for the fixed root hair case. After

48 hours have passed and the root hairs are fully developed the fixed and growing root hair

scenarios have identical uptake power. Hence, after 48 hours, the difference in cumulative

24

492

493

494

495

496

497

498

499

500

501

502

503

504

505

506

507

508

509

510

511

512

513

514

Page 25: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

uptake between the two scenarios is less than 1%. This suggests that the fixed root hair

provides a good approximation for nutrient uptake and that more detailed modelling of root

hair growth may not be necessary for timescales longer than 48 hours and shorter than the

root hair lifetime.

4. Conclusions

In this paper we have extended the image based modelling in (Keyes et al., 2013) to consider

large/infinite volume of soil around the single hairy root. The bulk soil properties are

captured from an X-ray CT image based geometry using the method of homogenization to

transform the imaged geometry into a homogeneous medium with an effective diffusion

constant based on the geometrical impedance offered by the soil. The bulk soil region is then

patched onto the rhizosphere using a boundary condition which relates the concentration at

the surface of the rhizosphere to the flux into the rhizosphere. The advantage to image based

modelling of this type is that it can be used to answer specific questions on the movement of

nutrient based on the soil geometry and root hair morphology. This information can then be

used to parameterise simplified models and gain understanding of rhizosphere processes.

The method is tested for two different soil water saturation values by considering bulk and

rhizosphere soil samples of different sizes which are increased until the effective transport

and uptake properties of the two regions are seen to converge. We found that the key

criterion for convergence of nutrient uptake simulations is that a sufficiently large radius of

soil about the root is considered. However, we emphasize that this is likely to be dependent

on root hair morphology, moisture content and soil type. Hence, convergence checks should

25

515

516

517

518

519

520

521

522

523

524

525

526

527

528

529

530

531

532

533

534

535

536

537

Page 26: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

be carried out on a smaller scale for different geometries. It is interesting to note that the

radius of the segment about the root needed for convergence to occur is dependent on the

saturation considered. Specifically we need to consider a larger region of soil about the root

for lower saturation values 1.4 mm of soil compared to the 1.1 mm of soil for the saturated

case. This observation is attributed to chemical effects present in the rhizosphere which

cause variation in the fluid properties (Gregory, 2006). In our CT scans these changes would

be picked up as a geometrical variation either in the fluid location or the contact angle at the

air water interface. Hence, in the fully saturated case we would not see these effects as the

specific water location is neglected. By dividing the CT image into regions ~r<0.2 ~rmax,

0.2 ~rmax<~r<0.4 ~rmax, 0.4 ~r max<~r<0.6 ~rmax, 0.6 ~rmax <~r<0.8 ~rmax and 0.8 ~rmax <~r<~rmax and

calculating the saturation in each case we see a noticeable decrease in saturation for

0.4 ~r max<~r<0.6 ~rmax, (S ≈ 0.5 compared to S>0.7 for all other regions). Therefore, in order to

capture the geometric impedance created by this region we require ~rb>0.6~rmax.

In agreement with the paper by (Keyes et al., 2013) we can observe that in both cases the root

hairs contribute less than the root surfaces to the nutrient uptake. However, this difference is

small and, in terms of order of magnitude, both the hairs and the roots contribute equally to

the nutrient uptake. A major advantage to this type of modelling is that these simulations can

be used to parameterise existing models which may be computationally less challenging, for

example the one in (Roose et al., 2001). In this case we see that the root hairs contribute

approximately an extra 30% to the root uptake, an important consideration which will

increase the predictive ability of the model developed in (Roose et al., 2001). Again, it may

be that in poorly saturated soils, i.e., those with much lower moisture content than the 33%,

there will be less water in contact with the root and the root hairs, which can penetrate into

26

538

539

540

541

542

543

544

545

546

547

548

549

550

551

552

553

554

555

556

557

558

559

560

561

Page 27: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

the smaller pores within the soil, may provide a more significant contribution. Further

investigation across a range of saturation values either based on imaging of samples at

different water content, or application of a spatially explicit moisture content model such as

the one developed in (Daly and Roose, 2015), would be needed to verify this.

Using the minimum value of ~h, ~rb and ~θ obtained for the two cases we have considered the

effect of pseudo root-hair growth on nutrient uptake neglecting the geometrical and

mechanical effects of root-hair growth. As the non-active root hairs are still geometrically

represented in the domain they may act to weakly impede the diffusion from the soil into the

root itself. The root hair growth modelling showed that, although there are differences in the

ratio of root and hair uptake rates towards the start of the growth period, i.e., two days.

However, on the long timescale, i.e., times larger than two days, the rate of uptake is

unaffected by the growing root hairs, see Figure 9. This suggests that more detailed

modelling of the root hair growth might not yield different results and the fixed root hair

approximation may provide a detailed enough picture for further investigation into the effects

of root hairs.

The method developed here provides a powerful framework to study properties of the

rhizosphere. We have tested the model for a specific geometry obtained from X-ray CT,

however we emphasise that the results obtained are suggestive of certain trends they will be

dependent on the precise soil type, water content and root hair morphology studied. Hence,

studying different root hair distributions and soil structures will allow a greater understanding

of how root hairs uptake nutrients to be developed. Further development of this method to

27

562

563

564

565

566

567

568

569

570

571

572

573

574

575

576

577

578

579

580

581

582

583

584

Page 28: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

couple a spatially explicit water model to the nutrient uptake will allow similar understanding

to be developed for water distribution and uptake.

5. Acknowledgements

The authors acknowledge the use of the IRIDIS High Performance Computing Facility, and

associated support services at the University of Southampton, in the completion of this work.

KRD is funded by BB/J000868/1 and by ERC consolidation grant 646809. SDK is funded by

and EPSRC Doctoral Prize fellowship, ERC consolidation grant 646809 and funding proved

by the Institute for Life Sciences at the University of Southampton. TR would like to

acknowledge the receipt of the following funding: Royal Society University Research

Fellowship, BBSRC grants BB/C518014, BB/I024283/1, BB/J000868/1, BB/J011460/1,

BB/L502625/1, BB/L026058/1, NERC grant NE/L000237/1, EPSRC grant EP/M020355/1

and by ERC consolidation grant 646809. All data supporting this study are openly available

from the University of Southampton repository at http://eprints.soton.ac.uk/id/eprint/384502

Data Accessibility

All reconstructed scan data will be available on request by emailing [email protected]

A. Diffusion in bulk soil

As in the rhizosphere transport is assumed to be by diffusion only

∂~C b

∂ ~t=~D~∇2~Cb ,

x∈Ωb , (A1)

28

585

586

587

588

589

590

591

592

593

594

595

596

597

598

599

600

601

602

603

604

605

Page 29: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

where ~Cb is the phosphate concentration in the soil solution for the bulk soil. We assume that

the concentration and concentration flux are continuous across the rhizosphere-bulk soil

boundary

~Cr=~Cb, x∈ Γrb , (A2)

~D n⋅~∇~C r=~D n ⋅~∇~Cb , x∈ Γrb , (A3)

and consider a linear reaction on the soil surface

~D n⋅~∇~Cb=−~ka~Cb+

~k d~Ca , x∈ Γsb , (A4)

∂~C a

∂ ~t=~ka

~Cb−~kd

~Ca ,x∈ Γsb , (A5)

and zero flux on the air water boundary

~D n⋅~∇~C b=0 , x∈ Γab , (A6)

In keeping with (Keyes et al., 2013) we non-dimensionalize with the following variables:

~Ca=[C a ]C a, ~Cr=[C r ]C r,

~Cb=[Cb ]Cb, ~t=[ t ] t , ~x= [ L ] x (for all lengths) and use [C r ]=[Cb ]=~K m,

[Ca ]=~Km

~ka~kd

, [ f ]=~D~Km

L and [ t ]=L2

~D to obtain

∂C b

∂ t=∇2Cb ,

x∈Ωb , (A7)

C r=C b , x∈ Γrb , (A8)

n ⋅∇C r=n ⋅∇Cb , x∈ Γrb , (A9)

n ⋅∇Cb=−δ1 (Cb−Ca ) , x∈ Γsb , (A10)

∂C a

∂ t=δ2 ( Cb−Ca ) ,

x∈ Γsb , (A11)

29

606

607

608

609

610

611

612

613

614

615

616

Page 30: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

n ⋅∇Cb=0 , x∈ Γab . (A12)

Here δ 1=~k a L/~D , δ 2=

~k d L2 /~D , λ=~Fm L~K m

~D, α=L~α and ~gr=

~gr L2

~D are summarized in Table 2

and L is a typical length scale of the domain we are interested in. Throughout this paper we

take L=10−3m, i.e. the average radius of the rice/wheat/corn fine roots.

Before we derive the boundary condition we consider the homogenization problem in the

bulk soil. The key assumption of the homogenization method is that the two length scales, Lx

and L, may be dealt with independently. Therefore, we can expand the vector gradient

operator as ∇=ϵ∇x+∇y, where ∇x and ∇ y are the vector gradient operators on the scale L x

and L respectively, ϵ=LLx

=10−2, corresponding to Lx=10−1 m. We also define the

concentrations in terms of an expansion of the form

Cb=Cb ,0+ϵ C b ,1+ϵ 2Cb , 2+ϵ3 Cb ,3+O (ϵ 4 ) (A13)

Ca=Ca ,0+ϵ C a ,1+ϵ 2Ca , 2+ϵ3 Ca ,3+O (ϵ 4 ) . (A14)

Physically Cb ,0 and Ca ,0 can be thought of as the average concentrations on the soil surface

and in the solution within the bulk soil domain. The remaining parameters Cb ,i and Ca ,i are

corrections to the average concentration which capture small scale variations within the soil

and the effects of the linear reactions on the soil surface. We rescale time to capture diffusion

over the distance L x (t x=ϵ2 t, δ 1=ϵ δ 1 and δ 2=δ 2) and obtain the scaled equations in the bulk

soil

30

617

618

619

620

621

622

623

624

625

626

627

628

629

630

631

632

633

634

Page 31: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

ϵ 2 ∂ Cb

∂ t x=∇2Cb ,

y∈Ωb , (A15)

n ⋅∇Cb=−ϵ δ1 (Cb−Ca ) , x∈ Γsb , (A16)

ϵ 2 ∂ Ca

∂ t x=δ

2(C b−Ca ) ,

x∈ Γsb , (A17)

n ⋅∇Cb=0 x∈ Γab , (A18)

periodic in y . (A19)

Substituting equations (A and (A14) into equation (A15) to (A19) and collecting terms in

ascending orders of ϵ we obtain a cascade of problems for Cb and Ca.

A.1 O(ϵ 0):

The equations at O(ϵ 0) are

∇ y2 Cb , 0=0 , y∈Ωb , (A20)

n ⋅∇ y Cb , 0=0 , x∈ Γsb , (A21)

δ 2 (Cb , 0−Ca ,0 )=0 , x∈ Γsb , (A22)

n ⋅∇Cb , 0=0 x∈ Γab , (A23)

periodic in y . (A24)

which have solution Ca ,0=Cb , 0=C 0(x). Hence, on the timescale of diffusion the linear

reactions have already happened and the surface concentration and the bulk concentration are

equal and the leading order concentration is independent of the microscale structure.

31

635

636

637

638

639

640

641

642

643

644

645

Page 32: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

A.2 O(ϵ 1):

Expanding to O(ϵ 1) and using the results from the O(ϵ 0) expansion we obtain

∇ y2 Cb , 1=0 , y∈Ωb , (A25)

n ⋅∇ y Cb , 1+ n⋅ ∇x Cb ,0=0 , x∈ Γsb , (A26)

δ 2 (Cb , 1−Ca , 1 )=0 , x∈ Γsb , (A27)

n ⋅∇ y Cb , 1+ n⋅ ∇x Cb ,0=0 x∈ Γab , (A28)

periodic in y. (A29)

This is the standard correction to a diffusion only problem, it has solution

Cb ,1=Ca ,1=C1 ( x )+C1(x , y ), where C1 ( x ) is a function of x only which will be defined at

higher order and

C1=∑p=1

3

χ p ∂xpC0 ,

x∈ Γ0 , (A30)

and χ p satisfies the cell problem

∇ y2 χ p=0 , y∈Ωb , (A31)

n ⋅(∇¿¿ y χ p+ ep)=0 ,¿ y∈Γ sb , (A32)

n ⋅(∇¿¿ y χ p+ ep)=0 ,¿ y∈Γ ab , (A33)

χ p periodic in y . (A34)

Equations (A31) to (A40) assume that the soil sample is geometrically periodic, i.e., it is

composed of regularly repeating units. In reality this is not the case so these equations cannot

32

646

647

648

649

650

651

652

653

654

655

656

Page 33: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

directly be applied. To overcome this we impose a mathematical periodicity by reflecting the

geometry along the three coordinate axes. Reflecting about the p-th coordinate axis using the

substitution y p →− y p in equations (A31) to (A40) we observe that, in order for the equations

to remain unchanged, we must have χ p →− χ p. Similarly by reflecting about the q-th

coordinate axis using the substitution yq →− yq, for q ≠ p in equations (A31) to (A40) we see

the equations remain unchanged. Hence we infer that χ p is odd when reflected about the p-th

coordinate axis and even otherwise. As a result we are able to solve a set of equations on the

original domain which respect these symmetries

∇ y2 χ p=0 , y∈Ωb , (A35)

n ⋅(∇¿¿ y χ p+ ep)=0 ,¿ y∈Γ s b , (A36)

n ⋅(∇¿¿ y χ p+ ep)=0 ,¿ y∈Γ ab , (A37)

χ p= 0, e p⋅ y=± 12

(A38)

eq ⋅∇ y χ p= 0, eq ⋅ y=± 12

, q ≠ p (A39)

A direct result of this analysis is that the dimensional volume used in equations (A35) to

(A39) represents one eighth of the total representative volume as the total volume can be

reconstructed by mirroring the result along the three axes.

A.3 O(ϵ 2):

Expanding to O(ϵ2) and using the previous solutions we obtain

33

657

658

659

660

661

662

663

664

665

666

667

668

669

670

671

Page 34: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

∂C b , 0

∂ t x=∇ y

2 Cb ,2+2∇x ⋅∇ y Cb ,1+∇x2 Cb , 0 ,

y∈Ωb , (A40)

n ⋅∇ y Cb , 2+ n⋅ ∇x Cb ,1=0 , x∈ Γsb , (A41)

∂Ca , 0

∂ t=δ

2(Cb ,2−Ca , 2) ,

x∈ Γsb , (A42)

n ⋅∇ y Cb , 2+ n⋅ ∇x Cb ,1=0 , x∈ Γab , (A43)

periodic in y . (A44)

Before we solve for C b ,2 and Ca ,2 we require that a solution exists. Integrating and using the

divergence theorem we obtain

||Ωb||∂C0

∂ t x=∇x ⋅∫

Ωb

[∇ y χ p⊗ e p+ I ]dy∇x C0 ,(A45)

where ¿|Ωb|∨¿∫Ωb

1dy. As in the O(ϵ 1) case we obtain a cell problem for Ca ,2 and Cb ,2.

Cb ,2=C2 (x )+C2(x , y ), where C2 ( x ) is an arbitrary function of x which will be defined at

higher order and

C2=∑p=1

3

χ p ∂xpC1+∑

p=1

3

∑q=1

3

ξ pq ∂x p∂xq

C0 .(A46)

Here ξ pq satisfies

¿|Ωb|∨[∇¿¿ y2 ξ pq+2 e p⋅ ∇ y χq+δ pq ]=e p⋅Deff eq ,¿ y∈Ωb , (A47)

n ⋅(∇¿¿ y ξ pq+e p χ q)=0 ,¿ y∈Γ sb , (A48)

34

672

673

674

675

676

677

678

679

680

681

Page 35: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

n ⋅(∇¿¿ y ξ pq+e p χ q)=0 ,¿ y∈Γ ab , (A49)

χ p periodic in y (A50)

and we have already solved the cell problem for χq.

A.4 O(ϵ 3):

In order to capture the effect of the linear reaction on the soil surface we expand to O(ϵ 3) to

obtain

∂Cb , 1

∂ t x=∇ y

2 C b ,3+2∇x ⋅ ∇y Cb ,2+∇x2Cb , 1 ,

y∈Ωb , (A51)

n ⋅∇ y Cb , 3+ n⋅ ∇x Cb ,2=−δ1 (Cb ,2−Ca , 2) , x∈ Γsb , (A52)

∂C a , 1

∂ t=δ

2(Cb ,3−Ca ,3 ) ,

x∈ Γsb , (A53)

n ⋅∇ y Cb , 3+ n⋅ ∇x Cb ,2=0 x∈ Γab , (A54)

periodic in y . (A55)

Applying the solvability condition and using the results from previous expansions we find

||Ωb||∂C1

∂ t x=∇x ⋅∫

Ωb

[∇y χ p⊗ ep+ I ] dy∇x C 1+¿(A56)

35

682

683

684

685

686

687

688

689

690

691

Page 36: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

∇x ⋅∫Ωb

∇ y ξ pq dy ∂x p∂xq

C0−||Γ||δ 1

δ 2

∂ C0

∂ t x,

where ¿|Γ|∨¿∫Γ

1 dy. By writing Cav=C0+ϵ C1 we can combine the results at O(ϵ 0) and

O(ϵ 1) to obtain

∂C av

∂ t x=∇x ⋅Deff ∇x Cav+ϵ ∇x ⋅H pq ∂x p

∂xqCav ,

(A57)

where we have neglected terms ∼O(ϵ 2). The effective parameters in equation (A57) are

Deff =∫Ωb

[∇ χ p⊗ ep+ I ]||Ωb||+¿|Γ|∨(δ 1/δ2)

dy(A58)

and

H pq=∫Ωb

∇ ξpq

||Ωb||+¿|Γ|∨(δ1/δ 2)dy .

(A59)

Note, we have added terms at O(ϵ2) to simplify the form of the final equation. The can be

formally calculated by expanding to O(ϵ 4). However, as we are only really interested in

knowing the approximate averaged behaviour of diffusion in the soil we neglect the final

term in equation (A57).

36

692

693

694

695

696

697

698

699

700

701

702

703

704

705

Page 37: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

B. Resupply boundary condition

We assume that equation (A57) is valid far from the root and root hair. We now consider an

infinite volume of soil surrounding the root and derive an approximate boundary “resupply”

condition at the interface between the Rhizosphere and Bulk soil regions to capture the influx

of nutrient from outside the explicit domain of interest. We assume that the main direction of

diffusion is towards the root. Hence, we can reduce the diffusion equation far from the root

to a one spatial dimension equation with an isotropic diffusion constant Deff . In the bulk soil

we consider the radial diffusion equation

∂Cb

∂t x=

Deff

r x

∂∂ r x

(r x

∂ Cb

∂ r x), (B1)

subject to the boundary conditions

Cb ( rb ,t x )=Cr (B2)

and

n ⋅∇x Cr=Deff

∂Cb

∂ rx.

(B3)

We add the condition that Cb ( ∞, t x )=C∞. We follow the approach of (Roose et al., 2001) and

solve equation (B1) in regions close to the root and far from the root. Matching these

solutions together at r x=rb we derive an expression for the flux into the rhizosphere. We

assume that, near the rhizosphere boundary, i.e., r x≪1 the concentration will approximately

be given by the steady state solution

37

706

707

708

709

710

711

712

713

714

715

716

717

718

719

720

721

722

723

Page 38: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

C r ≈C1+F rb ln (r x ) (B4)

where C1 is an arbitrary constant and F is the, as yet undefined, flux into the rhizosphere. Far

from the rhizosphere we use the similarity solution, see for example (King et al., 2003; Roose

et al., 2001),

Cb=C∞−B E1( r x2

4 D eff t x) (B5)

where E1 ( x )=∫x

∞ e− y

ydy is the exponential integral. By equating equations (B4) and (B5) at

r x=rb and assuming continuity of concentration on the rhizosphere-bulk soil boundary we

find

F=−2(C∞−C r)

rb [ γ−ln (4 D eff t x/rb2 ) ]

(B6)

where γ=0.57721 is the Euler–Mascheroni constant. We note that in deriving equation (B6)

we have implicitly assumed that F does not vary significantly with time. The logarithm in

the denominator of equation (B6) is singular at t x=0, hence, for numerical stability we

modify the equation slightly. This approach means that the boundary condition is not

accurate for small time. However, the timescale over which this occurs is short and the effect

on the uptake properties is small. Following the method of (Roose et al., 2001) we modify

equation (B6) such that F=0 at t x=0,

38

724

725

726

727

728

729

730

731

732

733

734

735

736

737

738

739

Page 39: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

F=−2(C∞−C r)

2+rb [γ−ln (4 D eff t x/rb2+eγ ) ]

(B7)

Equation (B7) is used to parameterize the bulk soil and feeds directly into the model for

nutrient movement in the rhizosphere in the main body of the paper.

2. References

Aravena, J. E., Berli, M., Ghezzehei, T. A. and Tyler, S. W. (2010). Effects of root-induced compaction on rhizosphere hydraulic properties-x-ray microtomography imaging and numerical simulations. Environmental Science & Technology 45, 425-431.

Aravena, J. E., Berli, M., Ruiz, S., Suárez, F., Ghezzehei, T. A. and Tyler, S. W. (2014). Quantifying coupled deformation and water flow in the rhizosphere using X-ray microtomography and numerical simulations. Plant and Soil 376, 95-110.

Barber, S. A. (1984). Soil Nutrient Bioavailability: A Mechanistic Approach: Wiley-Blackwell.

Daly, K. R., Mooney, S., Bennett, M., Crout, N., Roose, T. and Tracy, S. (2015). Assessing the influence of the rhizosphere on soil hydraulic properties using X-ray Computed Tomography and numerical modelling. Journal of experimental botany 66, 2305-2314.

Daly, K. R. and Roose, T. (2015). Homogenization of two fluid flow in porous media.

Datta, S., Kim, C. M., Pernas, M., Pires, N. D., Proust, H., Tam, T., Vijayakumar, P. and Dolan, L. (2011). Root hairs: development, growth and evolution at the plant-soil interface. Plant and Soil 346, 1-14.

Dexter, A. (1987). Compression of soil around roots. Plant and Soil 97, 401-406.Gahoonia, T. S. and Nielsen, N. E. (2004). Root traits as tools for creating

phosphorus efficient crop varieties. Plant and Soil 260, 47-57.George, T. S., Hawes, C., Newton, A. C., McKenzie, B. M., Hallett, P. D. and

Valentine, T. A. (2014). Field phenotyping and long-term platforms to characterise how crop genotypes interact with soil processes and the environment. Agronomy 4, 242-278.

Giesler, R. and Lundström, U. (1993). Soil solution chemistry: effects of bulking soil samples. Soil science society of America journal 57, 1283-1288.

Gregory, P. (2006). Roots, rhizosphere and soil: the route to a better understanding of soil science? European Journal of Soil Science 57, 2-12.

Hall, M., Frank, E., Holmes, G., Pfahringer, B., Reutemann, P. and Witten, I. H. (2009). The WEKA data mining software: an update. ACM SIGKDD explorations newsletter 11, 10-18.

Hallett, P. D., Karim, K. H., Glyn Bengough, A. and Otten, W. (2013). Biophysics of the vadose zone: From reality to model systems and back again. Vadose Zone Journal 12.

Hiltner, L. (1904). Über neuere Erfahrungen und Probleme auf dem Gebiete der Bodenbakteriologie unter besonderer Berücksichtigung der Gründüngung und Brache. Arbeiten der Deutschen Landwirtschaftlichen Gesellschaft 98, 59-78.

39

740

741

742

743

744

745746747748749750751752753754755756757758759760761762763764765766767768769770771772773774775776777778

Page 40: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

Hornung, U. (1997). Homogenization and porous media: Springer.Jasak, H., Jemcov, A. and Tukovic, Z. (2013). OpenFOAM: A C++ library for

complex physics simulations.Keyes, S. D., Daly, K. R., Gostling, N. J., Jones, D. L., Talboys, P., Pinzer, B. R.,

Boardman, R., Sinclair, I., Marchant, A. and Roose, T. (2013). High resolution synchrotron imaging of wheat root hairs growing in soil and image based modelling of phosphate uptake. New Phytologist 198, 1023-1029.

Keyes, S. D., Zygalakis, K. and Roose, T. (2015). Explicit structural modelling of root-hair and soil interactions at the micron-scale parameterized by synchrotron x-ray computed tomography. Submitted to Jounal of Experimental Botany 19/05/2015.

King, A. C., Billingham, J. and Otto, S. R. (2003). Differential equations: linear, nonlinear, ordinary, partial: Cambridge University Press.

Lucas, S. D. and Jones, D. L. (2006). Biodegradation of estrone and 17 β-estradiol in grassland soils amended with animal wastes. Soil Biology and Biochemistry 38, 2803-2815.

Ma, Z., Walk, T. C., Marcus, A. and Lynch, J. P. (2001). Morphological synergism in root hair length, density, initiation and geometry for phosphorus acquisition in Arabidopsis thaliana: a modeling approach. Plant and Soil 236, 221-235.

Murphy, J. and Riley, J. P. (1962). A modified single solution method for the determination of phosphate in natural waters. Analytica chimica acta 27, 31-36.

Nye, P. H. and Tinker, P. B. (1977). Solute movement in the soil-root system: Univ of California Press.

Pavliotis, G. and Stuart, A. (2008). Multiscale methods: averaging and homogenization: Springer Science & Business Media.

Roose, T., Fowler, A. and Darrah, P. (2001). A mathematical model of plant nutrient uptake. Journal of mathematical biology 42, 347-360.

Schnepf, A., Jones, D. and Roose, T. (2011). Modelling nutrient uptake by individual hyphae of arbuscular mycorrhizal fungi: temporal and spatial scales for an experimental design. Bulletin of mathematical biology 73, 2175-2200.

Tinker, P. B. and Nye, P. H. (2000). Solute movement in the rhizosphere, 2nd edition. USA: Oxford University Press.

Tracy, S. R., Daly, K. R., Sturrock, C. J., Crout, N. M. J., Mooney, S. J. and Roose, T. (2015). Three dimensional quantification of soil hydraulic properties using X-ray Computed Tomography and image based modelling Water Resources Research 51, 1006-1022.

Tracy, S. R., Roberts, J. A., Black, C. R., McNeill, A., Davidson, R. and Mooney, S. J. (2010). The X-factor: visualizing undisturbed root architecture in soils using X-ray computed tomography. Journal of experimental botany 61, 311-313.

Van Genuchten, M. T. (1980). A closed-form equation for predicting the hydraulic conductivity of unsaturated soils. Soil science society of America journal 44, 892-898.

Whalley, W. R., Riseley, B., Leeds‐Harrison, P. B., Bird, N. R., Leech, P. K. and Adderley, W. P. (2005). Structural differences between bulk and rhizosphere soil. European Journal of Soil Science 56, 353-360.

Wissuwa, M. and Ae, N. (2001). Further characterization of two QTLs that increase phosphorus uptake of rice (Oryza sativa L.) under phosphorus deficiency. Plant and Soil 237, 275-286.

Zygalakis, K., Kirk, G., Jones, D., Wissuwa, M. and Roose, T. (2011). A dual porosity model of nutrient uptake by root hairs. New Phytologist 192, 676-688.

Zygalakis, K. and Roose, T. (2012). A mathematical model for investigating the effect of cluster roots on plant nutrient uptake. The European Physical Journal-Special Topics 204, 103-118.

40

779780781782783784785786787788789790791792793794795796797798799800801802803804805806807808809810811812813814815816817818819820821822823824825826827828

Page 41: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

41

829

Page 42: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

Constant Value Units Description~D 10−9 m2 s−1 Diffusion constant~Km 6×10−3 molm−3 Uptake constant~Fm 5 ×10−9 mol m−2 s−1 Uptake rate~k d 2.569 ×10−4 s−1 Desorption rate~k a 3.73 ×10−9 m s−1 Adsorption rate~gr 5 ×10−9 m s−1 Root hair growth rate~α 0.2 ×106 m−1 Root hair cut off rate

Table 1: Physical parameters used in nutrient model.

Parameter Expression Values Descriptionδ 1

~ka L~D

3.73 ×10−3 Dimensionless flux coefficient

δ 2~kd L2

~D0.2569 Dimensionless desorption

rateλ ~Fm L

~Km~D

5/6 Dimensionless uptake rate

gr~gr L~D

5 ×10−3 Dimensionless root hair growth rate

α ~α L 200 Dimensionless root hair cut off rate

Table 2: Dimensionless parameters used in rescaled diffusion model.

42

830

831

832

Page 43: eprints.soton.ac.ukeprints.soton.ac.uk/384502/2/151127_Corrected.docx · Web viewImage based modeling of nutrient movement in and around the rhizosphere. Author list: Keith R. Daly1,

Figure captions

Figure 1: Image based geometry: (A) X-ray CT image of the roots, root hairs, soil and pore water. (B) shows one of the rhizosphere segments considered with size described by the angle θ, height h and radius r. (C) shows a cube of bulk soil of side length L.

Figure 2: Schematic of the different regions used in the mathematical formulation of the problem. The bulk and rhizosphere regions are highlighted with the modelling domains Ωr , Γ0 r, Γ0h, Γ sr, Γ ar and Γrb labeled for the

rhizosphere and Ωb, Γ sb and Γ ab labeled for the bulk soil.

Figure 3: Schematic for root hair growth. The root hair geometry remains unchanged throughout the simulation. As the simulation progresses the active portion of the root hair grows. The streamlines show an estimation of the nutrient transport as affected by the presence of the root hair. At t=0 the root hair is non-active and acts only to geometrically impede the transport of nutrient. At t=1 day the root hair has grown to half its final length and acts partly to take up nutrients and partly as a geometrical impedance. Finally at t=2 days the root hair is fully grown.

Figure 4: Effective diffusion coefficient as a function of L , the side length of the cube. The results are plotted for the saturated case, S=1, corresponding to the pore space being completely full of water and partially saturated, S=0.33, corresponding to 1/3 of the pore space being occupied by water. As L is increased the saturated and unsaturated effective diffusion properties are seen to converge to a steady value corresponding to the bulk diffusion coefficient.

Figure 5: Typical plots obtained from simulation for, (1A) to (1E). the saturated case and, (2A) to (2E), the partially saturated case. The five images in each case correspond to θ=0.2θmax=π /5 and h=0.2 hmax=0.36 mm for (A)

r=0.2 rmax=0.352mm, (B) r=0.4 rmax=0.704 mm, (C) r=0.6 rmax=1.056 mm, (D)

r=0.8 rmax=1.408mm and (E) r=1.0 rmax=1.76 mm. The streamlines show the effective diffusion paths with red corresponding to high nutrient concentration and blue corresponding to lower nutrient concentration.

Figure 6: Log-linear plot of flux into (A) the root hairs in the saturated case, (B) the roots in the saturated case, (C) the root hairs in the unsaturated case and (D) the roots in the unsaturated case. The lines show the average flux over a period of one week for the five different radii considered up to a maximum of r_max=1.76 mm. As r is increased the uptake profile is seen to converge. The figure insets show a zoomed in section of the same curve with error bars removed for clarity.

Figure 7: Total uptake into the root and root hair system averaged over all time points. Plot shows average uptake over the different radii considered with standard deviation error bars as a function of root surface area.

Figure 8: Root growth scenario for four different times: The top left shows the initial condition ( 0 hours), the top right shows the hair growth after 16 hours, the bottom left shows hair growth after 32 hours and the bottom right shows hair growth after 48 hours.

Figure 9: Flux at the root and hair surfaces for the root growth scenario in comparison to the fixed root hair scenario for the saturated and partially saturated geometries (logarithmic scale). The combined uptake of the root and hair is also plotted.

43

833

834835836

837838839

840841842843844

845846847848

849850851852853

854855856857858

859860

861862863

864865866