Potassium and calcium isotopic fractionation by plants (soybean [

Potassium and calcium isotopic fractionation by plants (soybean [...
1 downloads 0 Views 2MB Size
Subscriber access provided by Kaohsiung Medical University

Article

Potassium and calcium isotopic fractionation by plants (soybean [Glycine max], rice [Oryza sativa], and wheat [Triticum aestivum]) John N. Christensen, Liping Qin, Shaun T. Brown, and Donald J. DePaolo ACS Earth Space Chem., Just Accepted Manuscript • DOI: 10.1021/ acsearthspacechem.8b00035 • Publication Date (Web): 13 Jun 2018 Downloaded from http://pubs.acs.org on June 18, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

ACS Earth and Space Chemistry

Potassium and calcium isotopic fractionation by plants (soybean [Glycine max], rice [Oryza sativa], and wheat [Triticum aestivum]) John N. Christensen1* Liping Qin2 Shaun T. Brown1 Donald J. DePaolo1,3 1Earth

and Environmental Science Area, Energy Geosciences Division, Lawrence Berkeley National Laboratory, Berkeley, CA 94720. 2CAS Key Laboratory of Crust-Mantle Materials and Environment, University of Science and Technology of China, 96 Jinzhai RD., Hefei, Anhui 230026, China 3Department of Earth and Planetary Science, University of California, Berkeley, CA 94720. *Corresponding author

16 17

Keywords: Calcium isotopes; K isotopes; hydroponics; nutrient uptake and transport

18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36

Abstract We conducted hydroponic experiments growing soybean (Glycine max), rice (Oryza sativa) and wheat (Triticum aestivum) under K and Ca replete conditions to establish the degree of K isotopic fractionation by plants, and compare the isotopic fractionation of Ca and K. Each of the test plants displays fractionation relative to the growth solution favoring the light isotopes of K and Ca. The average δ41K of the roots from the three plant species were similar, and have an overall average of 0.55±0.24‰ 2s, while the overall average δ44Ca for roots is -0.67±0.44. For leaves, the overall average of δ41K is -0.97±0.4‰, compared to an overall average leaf δ44Ca of -0.83±0.09‰. In the case of the soybean plants, the lightest K and Ca occurs in the stems with average δ41K of -1.31±0.40‰ 2s and average δ44Ca of -1.20±0.19 ‰ 2s. We present a simple box model involving the relative fluxes of K and its isotopic fractionation that reproduces our K isotopic observations and suggests a fractionation of ~0.8‰ with K uptake from solution by roots. Directly comparing the per amu fractionation of K and Ca reveals an average factor of 2.05±0.50 2s greater fractionation of K isotopes which may reflect their different roles and behaviors in plants.

37

Potassium (K) and calcium (Ca) are key elements for plant physiology, being the

38

highest concentration cations in plants and, together with N and P, among the most

39

crucial elements for plant growth and productivity (1-3). K is centrally involved in a

40

wide range of crucial biophysical and biochemical processes in plants, including cell

41

osmotic pressure regulation, growth, regulation of photosynthesis, ion homeostasis,

42

control of cell membrane polarization, enzyme conformation/activation, water

Introduction

ACS Paragon Plus Environment

1

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 29

43

conservation and salt resilience (4, 5). Ca has its own crucial roles including as a

44

significant component in cellular walls, the maintenance of membrane integrity

45

against the passive transfer of H+, K+, Na+ and other monovalent cations, and

46

biochemical signaling (2, 6, 7). Due to these various essential roles, much research

47

has focused on the mechanisms of K and Ca uptake by and transport within plants

48

(8-15). For Ca and Mg, another key nutrient element, isotopic fractionation by

49

plants has been investigated both experimentally in the lab to elucidate processes of

50

acquisition and transport (16-19), but also in the field (e.g. 20-25) as monitors of the

51

local and global biogeochemical cycling (weathering) of these elements.

52

isotopic fractionation by plants of micronutrients such as Fe, Cu, Zn, and Mo have

53

also been measured (26-32). In the case of K it has been suggested that terrestrial

54

plants represent a significant reservoir for K, with as much as 40-70% or more of

55

the dissolved K in the world’s rivers coming from the decay of plant matter (33),

56

mass balance based on K/Rb), resulting from its crucial role as a plant nutrient.

57

Therefore, potassium isotopic systematics potentially provides a new tool for

58

tracking and quantifying nutrient cycling in ecological systems (e.g. boreal and

59

tropical forests); a proxy for global geochemical cycling; and a research avenue for

60

understanding the optimal fertilization for efficient agricultural food production. To

61

fulfill that potential, the degree of K isotopic fractionation by plants needs to be

62

established.

63

The

Two published studies do suggest that plants may fractionate K isotopes. 41K/39K

64

Deviations in the

65

favoring the light isotope have been measured by Li et al. (34) in several plant

66

materials (rice grains, chili peppers, tea leaves and Wolfberry fruit), but with no

67

measurement of the K sources. Likewise, in their broad survey of natural geologic

68

and organic materials Morgan et al. (35) found that a banana and a potato

69

purchased in a Scottish grocery had

70

respectively than seawater K, but again with no measurements of the K sources.

71

Here we use hydroponically grown plants, with isotopically characterized K and Ca

72

nutrient sources, to quantify and compare the isotopic fractionation of K and Ca by

ratio of ~1‰ from an in-house reference material

41K/39K

~0.4‰ lighter and ~0.1 ‰ heavier

ACS Paragon Plus Environment

2

Page 3 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

73

three species of vascular plants- soybean (Glycine max), rice (Oryza sativa) and

74

wheat (Triticum aestivum). These three plants were chosen to cover a selection of

75

important food crops; and each are C3 plants. We choose to compare K and Ca

76

isotopic fractionation for several reasons. First, the atomic weights of K and Ca are

77

relatively close so that mass fractionations can be easily compared. Secondly, K and

78

Ca have contrasting roles in plant physiology as described above, and have

79

contrasting valencies. Thirdly, there is a body of experimental data on Ca isotopic

80

fractionation by plants that we can compare our Ca isotopic results to (e.g. 18).

81

Methods

82

Soybean, rice and wheat were grown from seeds and cultivated in a large

83

hydroponic system at the U.C. Berkeley Oxford Facility greenhouse. The hydroponic

84

solution, a modified Hoagland’s solution replete with Ca nitrate (1.25 mM) and K

85

nitrate (2.5 mM, somewhat higher than the typical range of 0.1mM to 1mM in soil

86

porewaters (36), was replenished periodically to provide an isotopically constant

87

source of Ca and K during plant growth. Plants were harvested before full maturity

88

for analysis after growth periods of 2, 4, 6 and 8 weeks. Harvested plants were

89

divided broadly for soybeans into root, stem and leaf samples, and for rice and

90

wheat plants into samples of roots and leaves. The plant samples were rinsed, dried

91

and weighed, and then ashed before complete dissolution in high-purity nitric acid.

92

Aliquots were taken of the sample solutions for K and Ca isotopic analyses. K was

93

separated from the sample aliquots using AG50x8 cation resin and 1M HNO3 (37).

94

The K isotopic analyses were conducted on an IsoProbe MC-ICPMS at Lawrence

95

Berkeley National Lab (LBNL) using a sample-standard bracketing technique where

96

Ar-based mass interferences were removed (and peak tailing avoided) by the

97

introduction of Ne+H2 gas to the hexapole collision cell (38). Results are reported as

98

per mil deviations of the

99

using a spectroscopic concentration standard (ULTRA Scientific). We estimate that

100

our in-house standard on the Bulk Silicate Earth (BSE) scale of Wang and Jacobsen

101

(39) has an approximate δ41KBSE of +0.5‰. For Ca isotopic analysis, the sample

102

aliquots were spiked with a 42Ca-48Ca double spike before chemical separation using

41K/39K

ratio (δ41K) relative to an in-house K standard

ACS Paragon Plus Environment

3

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 29

103

Eichrom DGA resin eluted with 3N HNO3 and separated Ca collected with DI H2O (40,

104

41). The spiked Ca separates were then analyzed for isotopic composition using

105

thermal ionization mass spectrometry with a Triton multicollector instrument using

106

a multi-dynamic Faraday cup routine. The Ca isotopic results are reported as per

107

mil deviations of the

108

44Ca/40Ca

109

is -1.0±0.1 2s. Further details of the K and Ca separation and isotopic analyses are

110

presented in the Supplement.

111

Results

44Ca/40Ca

ratio (δ44Ca) from the Bulk Silicate Earth (BSE)

of 0.0212035 (42). On this scale our long-term average δ44Ca of SRM915A

112

The K isotopic composition of the K-nitrate used in the hydroponic solutions is

113

indistinguishable from the in-house standard (Table 1). Likewise, the Ca isotopic

114

composition of the Ca nitrate is indistinguishable from BSE (Table 1). For both K

115

and Ca isotopic compositions, each of the test plants show fractionation favoring the

116

light isotope (Figure 1). No resolvable trends with growth time in K or Ca isotopic

117

composition are observed, which indicates that replenishment of the hydroponic

118

solution provided an effectively infinite K and Ca source reservoir during plant

119

growth, and that the size of the plants had no discernable effect on the magnitude of

120

the isotopic fractionation for each species. The greatest degree of K and Ca isotopic

121

fractionation is seen in the soybean plants, followed by wheat and then rice. The

122

δ41K analyses of plant roots represents K acquisition, and indicates a fractionation of

123

-0.65±0.22‰ for soybean (average±2s, n=4), -0.40±0.17‰ for rice (average±2s,

124

n=4) and -0.74±0.36 (average±2s, n=3) with an overall average δ41K of -

125

0.55±0.24‰ 2s. For Ca, the roots gave δ44Ca of -0.50±0.26‰ for soybean

126

(average±2s, n=4), -0.67±0.4 for rice (average±2s, n=3), -0.82±0.24 (average±2s,

127

n=3) for wheat with an overall average δ44Ca of -0.67±0.44. The average K isotopic

128

compositions of leaves are -1.216±0.095‰ (n= 4) for soybean, -0.62±0.18‰ (n=4)

129

for rice, and -1.11±0.46‰ (n=4) for wheat with an overall average of -0.97±0.4‰,

130

compared to an overall average leaf δ44Ca of -0.83±0.09‰. The root-leaf difference

131

in δ41K for each of the plants is small with the average difference for soybean of

132

0.57±0.37‰, for rice of 0.26±0.28‰ and for wheat 0.59±0.3‰. The average root-

ACS Paragon Plus Environment

4

Page 5 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

133

leaf δ41K difference for all the plants is resolvable from zero at 0.47±0.16‰

134

however the average root-leaf difference in δ44Ca of 0.03±0.25‰ is essentially zero.

135

For the soybean plants, where stems were analyzed in addition to roots and leaves,

136

the lightest K occurs in the stems (δ41K ave. =-1.31±0.40‰ 2s) as does the lightest

137

Ca (δ44Ca ave = -1.20±0.19 ‰ 2s), with average root-stem differences of

138

0.64±0.26 ‰ for δ41K and 0.70±0.23 % for δ44Ca. A similar root-stem difference in

139

δ44Ca of 0.45±0.09 ‰ has been observed in experiments with French beans under

140

similar growth conditions (18). A small or zero root-leaf difference in δ44Ca is also

141

consistent with the observations in Cobert et al. (18).

142

The overall pattern of isotopic fractionation of K and Ca is similar for each of the

143

plants, but there is a difference in the relative size of the fractionation. Since

144

different mass differences are involved in the measured isotopic compositions of K

145

(41K &

146

directly compare the isotopic fractionation of these two elements by plants it is

147

useful to compare them on a per amu basis (Table 3). The average ratio of the per

148

amu fractionation of K isotopes to Ca isotopes is 2.05±0.50 suggesting an

149

approximate factor of 2 larger discrimination by plants of the light K isotope

150

compared to the light Ca isotope, though they have similar atomic masses.

151 152

Discussion K and Ca transport flux model and fractionation factors

39K,

a difference of 2 amu) and Ca (44Ca &

40Ca,

a difference of 4 amu), to

153

We have measured differences in isotopic ratios between different parts of the

154

plants and between the plants and the source K and Ca in the hydroponic solution

155

(our analog for “soil porewater”). We would, however, also like to say something

156

about the isotopic fractionation that accompanies the movement of K and Ca

157

between “soil porewter” and roots and then within the plant. This fractionation may

158

not be simply the difference in isotopic ratio, because each part of the plant has

159

fluxes of K and Ca entering and exiting. To address this issue we have constructed a

160

fairly simple box model that treats each of the major parts of the plant (roots, stem,

161

leaves) as a separate reservoir and evaluate how K and Ca isotopes would need to

162

be differentially transported to explain the observations.

ACS Paragon Plus Environment

5

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 29

163

The model and the associated equations (Figure 2) are explained in more detail

164

in the Supplement. Isotopic fractionation is assumed to occur during transfer of K

165

and Ca between reservoirs. The isotopic fractionation factors (ε) represent the

166

difference in δ41K (or δ44Ca) between the K (or Ca) transported out of a reservoir,

167

and the average value in the reservoir. The accumulation rate of K in each reservoir

168

is assumed to be proportional to the (mass) growth rate of that reservoir. For

169

example, the growth rate of roots is proportional to F1+F4-F2, and the growth rate of

170

stems is proportional to F2-F3. The ratio of the growth rate of root to the growth

171

rate of shoots (=stem + leaves) is [F1/(F2-F4)]-1. The proportionality between K

172

content and mass is consistent with the generally constant concentration of K in

173

plant biomass, although what differences there are would need to be factored in for

174

relating the model flux values to mass changes.

175

The observed difference between root K and hydroponic solution K (“soil

176

porewater” K referred to as “soil”) is in the range of ∆root/soil = -0.5 to -0.9‰. This

177

observed difference is a function of various K fluxes (Fi) and isotopic fractionation

178

factors (εi) as shown in Figure 2. The model and data do not allow for much latitude

179

with regard to the fractionation factors, but we work through the analysis to show

180

how the model can be used to deduce the εi values. We begin with the ∆ values

181

(difference between δ41K values) between leaf and stem, or between stem and root.

182

For example the typical values of ∆stem/root are -0.5 to -0.8 ‰ (Table 2). Since leaves

183

are typically heavier than stems, and can be affected by recycling of K back to the

184

roots, it may be reasonable to set ε3 = 0 ‰ as a first guess, which then requires that

185

ε2 ≈ -0.5 to -0.8 ‰ (fractionation between the root-to-stem flux and roots; Figure 2,

186

eqn. B). Because the bean leaves typically have δ41K that is similar to or slightly

187

higher than that of stems, it suggests that ε4 has a small negative value, which we

188

estimate as -0.2 ‰ (or slightly larger because F3/F4 < 1 if there is leaf growth;

189

Figure 2, eqn. A).

190

Given that we have values for ε2 (-0.6 ‰) and ε4 (-0.2 ‰) we can substitute

191

them into the equation for the (measured) effective root/soil fractionation (Figure 2,

192

eqn. C):

ACS Paragon Plus Environment

6

Page 7 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

193

∆ root/soil = 194

F1 F F ε1 − 2 ε 2 + 4 (ε 4 + ∆leaf /soil ) F1 + F4 F1 + F4 F1 + F4

(1)

195 196

We have measured ∆root/soil to be -0.5 to -0.8 ‰, so by the above reasoning we now

197

have values for all of the parameters except ε1 and the fluxes.. To go farther we need

198

approximate values for the K flux ratios, and we use F1:F2:F3:F4 = 1: 0.8 : 0.6 : 0.4

199

based on published experiments on lupin (43). With these values the stems and

200

leaves are growing at the same rate, and 40% of the K being taken up by the plant is

201

being returned to the roots from the leaves. The root/shoot mass ratio would be 1.5,

202

which is a typical value (44). Substituting into equation (1), and estimating ∆leaf/soil =

203

-1.2 ‰ (Table 2), gives ε1 = -0.7 to -1.0 ‰ (fractionation between root and soil pore

204

water). Using this value and the flux ratios given above we can then deduce a self-

205

consistent set of isotopic ∆ values:

206

∆root/soil = -0.6 ‰; ∆stem/root = -0.6 ‰; ∆leaf/stem = +0.1 ‰ and ∆leaf/soil = -1.1 ‰.

207

Thus, given a plausible set of fluxes, and the fractionation factors (ε’s) deduced

208

above, we can reproduce the observed general pattern and magnitudes of K isotopic

209

distribution. Our analysis suggests that the isotopic discrimination associated with K

210

uptake from the hydroponic solution is the largest (ε1 = -0.7 to -1.0 ‰), the

211

discrimination associated with transfer of K from root to stem is the next largest (ε2

212

≈-0.6 ‰), and the other two discrimination factors are small (0 to -0.2 ‰).For the

213

example given, if the hydroponic solution has δ41K = 0, then the δ41K values for the

214

roots, stems, and leaves would be -0.6, -1.2, and -1.1 ‰ respectively, similar to

215

those observed (Table 2). As a check, the bulk plant ∆41K value relative to the

216

solution muct be equal to ε1. We do not have measurements of the relative masses of

217

the plant parts, but it is evident from the data that the bulk plant has a δ41K

218

somewhat lower than the roots, and hence is likely to be in the same range as

219

inferred for ε1 (i.e. ∆41Kplant-solution ≈ -0.7 to -1.0 )

ACS Paragon Plus Environment

7

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 29

220

The model can help evaluate isotopic fractionation in the plants under different

221

growth conditions, because the observed isotopic differences between plant

222

components depend on the flux ratios, even if the fractionation factors are constant.

223

For example, if roots are not growing significantly in comparison to shoots (F2/F1 ->

224

1), then the observed root/soil fractionation becomes smaller even though the

225

fractionation factor (ε1) remains the same. If roots are growing much faster than

226

shoots, then the observed root/soil fractionation becomes larger. These flux ratios

227

are likely to vary over the lifetime of the plant, and also depend on soil conditions

228

and nutrient availability.

229

The transport model for Ca is somewhat simpler than that for K in that there is

230

virtually zero return of Ca to the roots from leaves (43).

However, there is a

231

complication in that for Ca the leaves are more consistently higher in δ44Ca than

232

stems compared to the pattern for δ41K. In their Ca-replete experiments Cobert et al.

233

(18) also found stems to have lower δ44Ca than roots or leaves. The only way to

234

accommodate these observations in the context of our model is to have the

235

parameter ε3 have a positive value for Ca isotopes. This would tend to make the

236

stems lighter than both the leaves and roots. The adsorption of light Ca (low δ44Ca)

237

into cell walls lining xylem pathways has been suggested as the mechanism for

238

isotopic fractionation between roots and shoots (45, 18, 19), and would be

239

consistent with a positive ε3. This light Ca loss during transport would cause the Ca

240

carried in the xylem to become progressively enriched in heavy isotopes, making the

241

leaves generally higher in δ44Ca than stems. Excess Ca sequestered as Ca-oxalate

242

nodules could enhance this fractionation, as Ca-oxalate derived from leaves has

243

been shown to have lower δ44Ca than the free-Ca in leaves (18).

244

Mechanisms of K and Ca transport in plants

245

The model described above is non-specific with regard to the transport

246

mechanisms and source of isotopic fractionation. A number of different mechanisms

247

have been called upon to explain isotopic fractionation within plants (17-19, 28-32,

248

45,46) including adsorption or ligand binding, transport processes across cellular

249

membranes some of which involve transporter proteins, and diffusion within plant

ACS Paragon Plus Environment

8

Page 9 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

250

fluids involving differences in diffusivity between isotopes. This last mechanism,

251

diffusion within the vascular fluids of plants, has been proposed for Zn by Moynier

252

et al. 2009 (31), which has been also taken up as on mechanism for Cu isotopic

253

fractionation in plants (28). Moynier et al. (31) in considering this mechanism

254

calculate over a 10cm length scale the fractionation of Zn isotopes by diffusion and

255

derive a value for the ratio of the diffusivities of 66Zn to 64Zn within a factor of 10 of

256

a published value for Zn diffusion in water and conclude that this mechanism could

257

explain, at least in part, their observation of increasing Zn isotopic fractionation

258

with leaf height. However their diffusion model assumes a stagnant fluid, and does

259

not take into account fluid movement within the plant. To resolve whether diffusive

260

isotopic fractionation can be a significant mechanism, we calculate the Péclet (Pe)

261

number (Eq. 2), a dimensionless number that compares the relative importance of

262

fluid advective transport to diffusion in the fluid:

263

Pe = vL/D

Eq. 2

264

where v is the fluid velocity, L is the length scale and D is the diffusivity. For Pe < 1

265

diffusive transport dominates, for Pe > 1 advective transport dominates. Using a

266

length scale of 10cm, a fluid velocity in the xylem of between 0.25 mm/sec (night

267

conditions) and 0.4 mm/sec (day conditions) from measurements in castor bean

268

plants (47) and a diffusivity of K in water of 1.8x10-9 m2/sec (48) yields a Pe number

269

of on the order of 104, indicating that diffusive effects within xlyem and phloem are

270

completely negligible. This is likely true for other nutrient elements in plants, since

271

they likely have diffusivities that are within an order of magnitude of 10-9 m2/sec.

272

Plant cellular cytoplasmic resting concentrations of Ca2+ are on the order of

273

100nM, a factor ~106 less than cytoplastic K+ concentrations, and can rise to ~1 µM

274

in its cellular role in molecular signaling (2, 6, 36). This requirement of extremely

275

low Ca2+ concentration in the cytoplasm is a result of its tendency to form insoluble

276

salts with sulfates, phosphates and other compounds that would be detrimental for

277

the cell. Instead, Ca2+ is stored in cell organelles such as vacuoles at ~1mM levels, or

278

in the apoplast. Unlike K+, Ca2+ has key structural roles in providing binding in

279

cellular membranes, as well as in cellular walls and other rigid structures of plants.

ACS Paragon Plus Environment

9

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 29

280

Based on this structural role for Ca2+ and Ca experimental isotopic observations,

281

Cobert et al. (18) and Schmitt et al. (19) call on a solid/fluid ion exchange process

282

(akin to ion chromatography) to explain plant Ca isotopic fractionation, both during

283

Ca2+ uptake by plant roots, as well as during apoplastic and xylem transport, and

284

precipitation in leaves of Ca-oxalate (this latter process is emphasized by both (19)

285

and by Moynier and Fujii (49) for root-leaf fractionation). In contrast to Ca2+, the

286

bulk of which in plants participates in solid/liquid exchange processes, K+

287

participates mostly as an ion in plants (4).

288

In order to maintain the ~100 mM cellular cytoplasm K+ concentration, K+ must

289

be driven up a concentration gradient from the lower K+ concentration external

290

solution, either from the soil solution or from extracellular fluids (4). For K

291

transport into plant cells, two broad mechanisms operate: a high affinity K transport

292

system (HATS) under low K+ external concentration ([K+]1mM),

294

reflective of its key and multifaceted role in plant nutrition (9). The LATS uses

295

highly selective K-specific ion channels across the cellular membrane, while the

296

HATS uses transporter proteins and synporter systems. For our experiments with

297

[K+] = 2.5mM, the LATS is likely the dominant mode of K+ influx from the hydroponic

298

solution to root cells and throughout the plant, the energetically intensive HATS

299

mode being saturated and down-regulated. After uptake by root cells, K+ follows a

300

variety of routes (symplastic and apoplastic) to the xylem fluid for distribution

301

throughout the plant, delivery to symplastic and apoplastic pathways for trans-

302

membrane uptake by cellular tissues (11).

303

through the xylem is hydraulically driven by transpiration from the leaves, and root

304

pressure arising from osmotic processes (50). Thus the most likely processes

305

affecting K+ isotopic fractionation are the various means of transport across cellular

306

membranes.

Transport of solute-bearing fluids

307

K+ specific channels for transport across cellular membranes (favored in LATS)

308

can be highly selective; Na+ for example can be selected against by a factor of 10,000

309

compared to K+ (50), that is for every 10,000 K atoms entering a K+ specific channel

ACS Paragon Plus Environment

10

Page 11 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

310

only one Na ion does. It has been suggested that the 25% difference in ionic radius

311

between Na+ and K+ is a key factor in this selectivity (51, 52). Figure 3 shows a plot

312

of estimated selectivities relative to K+ as a function of ionic radius. With ionic radii

313

differences of 9% for Rb+ and 19% for Cs+, selectivity against those ions is also fairly

314

significant (51, 53), further suggesting a partial role for ionic radius in the selectivity

315

of K+ channels.

316

relationship to K+ isotopic fractionation. Ionic radius is affected by isotopic mass,

317

with smaller radii trending with increasing nuclear mass (i.e. number of neutrons)

318

for an element. For instance, 22Ne is ~0.2% smaller than 20Ne (54) (Figure 4). Using

319

the difference in ionic radius between Na+ and K+ to scale the selectivity against Na+

320

by K+ specific channels we extrapolated to a size difference between 41K and 39K ions

321

of 0.0035% required to produce a 1‰ selectivity against

322

would be sufficient to account for our K isotopic observations. This estimated

323

radius difference between 41K and 39K is consistent with the trend shown in Figure 4.

324

The K+ specific channel may be optimized to the size of

325

isotope of K (39K is about 93.3% while 41K is about 6.7% of K).

This apparent sensitivity to ionic radius suggests a possible

41K

39K+,

relative to

39K,

that

the more common

326

Another possible effect to consider with transport across cellular membranes is

327

that ions would need to be desolvated before their passage. Hoffman et al. (55)

328

suggest through molecular dynamics simulations, that water exchange rates for

329

solvated ions has a rate dependence on the mass of the solvated ion such that the

330

lighter isotope is favored in desolvation, resulting in a predicted ~2‰ fractionation

331

in

332

feature of K+ specific channels lies in the process of desolvation/solvation of K+

333

during transmembrane transport. Solvated K+ enters the channel where the K+ is

334

handed off to eight protein bonded oxygen atoms replacing the role of the solvating

335

H2O molecules, minimizing the energy required for desolvation, with resolvation

336

occurring with the exit of K+ from the channel (56). Therefore we suggest that

337

isotopic fractionation due to dehydration of K+ accompanying K+ channel transport

338

may be minimal. Instead we hypothesize that their size selectivity as argued above

339

is the main source of fractionation of K isotopes.

41K/39K

between a precipitating solid and the solution. However, a further

ACS Paragon Plus Environment

11

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 29

340

We propose that the HATS, operative under low external K+, would fractionate K

341

isotopes less that the LATS, which in large part depends on highly selective K+ ion

342

channels. This suggestion is supported by experiments involving cells of a marine

343

diatom where under low external Zn concentration (HA-high-affinity transport), the

344

observed isotopic fractionation of Zn is four times less than under high Zn external

345

concentration (LA- low affinity transport) (57). A difference between HA vs. LA

346

transport has also been called upon by Deng et al. 2014 (58) to explain differences

347

in Zn isotopic fractionation between Zn nonaccumulating and Zn hyperacumulating

348

plants. For isotopic fractionation of K in plants, a test of this would be examining K

349

isotopic fractionation as a function of external solution K+ concentration. Another

350

possible test would be experiments involving genetically modified plants that are

351

capable of only either HATS or LATS K+ transport to look for contrasts in K isotopic

352

fractionation.

353 354

Conclusions We have shown through hydroponic growth experiments that soybean, wheat

355

and rice can fractionate the isotopes of potassium relative to the nutrient source,

356

with all portions of the plant lighter than the source K. While the pattern of root to

357

leaf K isotope fractionation is similar to that of the isotopic fractionation Ca, the

358

magnitude of the fractionation on a per amu basis is about 2 times greater for K than

359

for Ca, and may be related to different mechanisms of fractionation- in the case of Ca

360

mainly association with solid/liquid exchange (18) and precipitation of oxalates in

361

leaves (as proposed by (19, 49)) versus K isotopic fractionation associated mainly

362

with trans-membrane transport.

363

The box-model proposed here may be a useful framework for using K and Ca

364

isotopes in plants to constrain the relative transport fluxes and mechanisms for

365

these elements, and how they change with nutrient availability, light levels, plant

366

maturity and other factors.

367

Since K is a significant nutrient for plants, and is often added as fertilizer in

368

agriculture (60, 61), there may be advantage in being able to isotopically distinguish

369

between resident soil K, the K added as fertilizer, the K in the plants, and how it’s

ACS Paragon Plus Environment

12

Page 13 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

370

recycled back (or lost) in fallow periods. Drought and salt stress is a significant

371

challenge in agriculture, and understanding the roles and mechanism of K and Ca in

372

the response of plants to stress is an active area of research (61). From our

373

discussion, as Ca and K appear to be fractionated by different processes in plants,

374

the comparisons between their isotopic behaviors with variation of nutrient

375

availability and salt loading may prove illuminating. In particular, conducting

376

experiments with mutant plant varieties with different trans-membrane transport

377

processes shut off may provide insights into the operation of K+ transport channels

378

and their ion selectivity.

379

Our results also bear on the global geochemical cycling of K. Seawater K is nearly

380

0.6 ‰ heavier than BSE (39, 24), and based on K isotopic analyses of porewater

381

from oceanic sediments and modeling of that data Santiago Ramos et al. 2018 (62)

382

suggest that weathering of sediments on the seafloor is a sink for

383

seawater K to a heavy δ41K value. From our experiments it would appear that plants

384

would be either lighter than the BSE, or about the BSE value depending on K source.

385

With up to 40-70% of the dissolve K load of rivers coming from plant decay (33), the

386

light K of plants should be a driver of seawater K toward lighter values, with the best

387

test being a K isotopic survey of major rivers. This emphasizes the need identified

388

by Santiago Ramos et al. (62) for an oceanic sediment sink for light K.

389 390

Acknowledgements

39K,

driving

391 392 393 394 395 396 397

This research was supported by from the US Department of Energy, Office of Science, Office of Basic Energy Sciences under Award No. DE-AC02-05CH11231 to Lawrence Berkeley National Laboratory. We wish to thank Barbara Rotz of the Oxford Greenhouse for her assistance in setting up and tending the hydroponic experiments. We also thank three anonymous reviewers whose comments helped us improve the manuscript.

398 399

References 1. Dreyer, I. (2014) Potassium (K+) in plants. J. Plant Physiol., 171:655.

400

2. White, P.J., Broadly, M.R. (2003) Calcium in plants. Annals of Botany, 92:487-511.

401 402

3. Clarkson, D.T., Hanson, J.B. (1980) The mineral nutrition of higher plants. Ann. Rev. Plant Physiol., 31:239-298.

ACS Paragon Plus Environment

13

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 29

403 404

4. White, P.J., Karley, A.J. (2010) Potassium. In Cell Biology of Metals and Nutrients. R. Hell and R.-R. Mendel eds., pp.199-224. Springer-Verlag, Berlin

405 406

5. Wang, Y., Wu, W.-H. (2017) Regulation of potassium transport and signaling in plants. Current Opinion Plant Bio., 39:123-128.

407 408

6. Batisticˇ, O., Kudla, J. (2010) Calcium: Not just another ion. In Cell Biology of Metals and Nutrients. R. Hell and R.-R. Mendel eds., pp.17-54. Springer-Verlag, Berlin

409 410

7. Hirschi, K.D. (2004) The calcium conundrum. Both versatile nutrient and specific signal. Plant Physiol., 136:2438-2442.

411 412

8. Lebaudy, A., Véry, A.-A., Sentenac, H. (2007) K+ channel activity in plants: Genes, regulations and functions. FEBS Lttrs., 581:2357-2366.

413 414

9. Britto D.T., Kronzucker H.J. (2008) Cellular mechanisms of potassium transport in plants. Physiol. Plant.,133:637–650.

415 416 417

10. Véry, A.-A., Nieves-Cordones, M., Daly, M., Khan, I., Fizames, C., Sentenac, H. (2014) Molecular biology of K+ transport across the plant cell membrane: What do we learn from comparison between plant species ? J. Plant Physiol., 171:748-769.

418 419

11. Ahmad, I., Maathuis, F.J.M. (2014) Cellular and tissue distribution of potassium; physiological relevance, mechanisms and regulation. J. Plant. Physiol., 171:708-714.

420 421 422

12. Nieves-Cordones, M., Alemá, F., Martínez, V., Rubio, F. (2014) K+ uptake in plant roots. The systems involved, their regulation and parallels in other organisms. J. Plant. Physiol., 171:688-695.

423 424

13. Jammes, F., Hu, H.-C., Villiers, F., Bouten, R., Kwak, J.M. (2011) Calciumpermeable channels in plant cells. FEBS J., 278: 4262-4276.

425 426 427

14. Gilliham, M., Dayod, M., Hocking, B.J., Xu, B., Conn, S.J., Kaiser, B.N., Leigh, R.A., Tyerman, S.D. (2011) Calcium delivery and storage in plant leaves: exploring the link with water flow. J. Exper. Botany, 62(7):2233-2250.

428 429 430

15. Tang, R.-J. and Luan, S. (2017) Regulation of calcium and magnesium homeostasis in plants: from tranporters to signaling network. Current Opinion Plant Bio., 39:97-105.

431 432

16. Black J.B., Epstein, E., Rains, W.D., Yin, Q.-Z., Casey, W.H. (2008) Magnesiumisotope fractionation during plant growth. Environ. Sci. Tecnol., 42:7831-7836.

433 434 435

17. Bolou-Bi, E.B., Poszwa, A., Leyval, C., Vigier N. (2010) Experimental determination of magnesium isotope fractionation during higher plant growth. Geochim. Cosmochim. Acta, 74:2523-2537.

436 437 438

18. Cobert, F., Schmitt, A.-D., Bourgeade, P., Labolle, F., Badot, P.-M., Chabaux, F., Stille, P. (2011) Experimental identification of Ca isotope fractionations in higher plants. Geochim. Cosmochim. Acta, 75:5467-5482.

439 440 441

19. Schmitt, A.D., Cobert, F., Bourgeade, P., Ertlen, D., Labolle, F., Gangloff, S., Badot, P.-M., Chabaus, F., Stille, P. (2013) Calcium isotope fractionation during plant growth under a limited nutrient supply. Geochim. Cosmochim. Acta, 110:70-83.

ACS Paragon Plus Environment

14

Page 15 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

442 443

20. Holmden, C., Bélanger, N. (2010) Ca isotope cycling in a forested ecosystem. Geochim. Cosmochim. Acta, 74:995-1015.

444 445

21. Hindshaw, R.S., Reynolds, B.C., Wiederhold, J.G., Kiczka, M., Kretzschmar, R., Bourdon, B. (2013) Biogeochemistry, 112:373-388.

446 447

22. Bullen, T., Chadwick, O. (2016) Ca, Sr and Ba stable isotopes reveal the fate of soil nutrients along a tropical climosequence in Hawaii. Chem. Geol., 422:25-45.

448 449 450 451

23. Schmitt, A.-D., Gangloff, S., Labolle, F., Chabaux, F., Stille, P. (2017) Calcium biogeochemical cycle at the beech tree-soil solution interface from the Strengbach CZO (NE France): insights from stable Ca and radiogenic Sr isotopes. Geochim. Cosmochim. Acta, 213:91-109.

452 453 454

24. Bolou-Bi, B.E., Vigier, N., Poszwa, A., Boudot, J.P., Dambrine, E. (2012) Effects of biogeochemical processes on magnesium isotope variations in a forested catchment in the Vosges Mountains (France). Geochim. Cosmochim. Acta, 87:341-355.

455 456 457

25. Uhlig, D., Schuessler, J.A., Bouchez, J., Dixon, J.L., von Blanckenburg, F. (2017) Quantifying nutrient uptake as driver of rock weathering in forest ecosystems by magnesium isotopes. Biogeosci.s, 14:3111-3128.

458 459

26. Guelke, M., Von Blanckenburg, F. (2007) Fractionation of stable iron isotopes in higher plants. Environ. Sci. Technol., 41:1896-1901.

460 461 462

27. Kiczka, M., Wiederhold, J.G., Kraemer, S.M., Bourdon, B., Kretzschmar, R. (2010) Iron isotope fractionation during Fe uptake and translocation in Alpine plants. Environ. Sci. Technol., 44:6144-6150

463 464

28. Weinstein, C., Moynier, F., Wang, K., Paniello, R., Foriel, J., Catalano, J., Pichat, S. (2011) Isotopic fractionation of Cu in plants. Chem. Geol., 286:266-271.

465 466 467 468

29. Jouvin, D., Weiss, D.J., Mason, T.F.M., Bravin, M.N., Louvat, P., Zhao, F., Ferec, F., Hinsinger, P., Benedetti, M.F. (2012) Stable isotope of Cu and Zn in higher plants: Evidence for Cu reduction at the root surface and two conceptual models for isotopic fractionation processes. Environ. Sci. Technol., 46:2652-2660.

469 470 471

30. Weiss, D.J., Mason, T.F.D., Zhao, F.T., Kirk, G.J.D., Coles, B.J., Horstwoo, M.S.A. (2005) Isotopic discrimination of zinc in higher plants. New Phytologist, 165(3):703710.

472 473

31. Moynier, F., Pichat, S., Pons, M.-L., Fike, D., Balter, V., Albarède, F. (2009) Isotopic fractionation and transport mechanism of Zn in plants. Chem. Geol., 267:125-130.

474 475

32. Malinovsky, D., Kashulin, N.A. (2018) Molybdenum isotope fractionation in plants measured by MC-ICPMS. Anal. Methods, 10:131-137.

476 477

33. Chaudhuri, S., Clauer, N., Semhi, K. (2007) Plant decay as a major control of river dissolved potassium: A first estimate. Chemical Geol., 243:178-190.

478 479 480

34. Li, W., Beard, B.L., Li, S. (2016) Precise measurement of stable potassium isotope ratios using a single focusing collision cell multi-collector ICP-MS. J. Anal. At. Spectrom., 31:1023-1029.

ACS Paragon Plus Environment

15

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 29

481 482 483

35. Morgan, L.E., Santiago Ramos, D.P., Davidheiser-Kroll, B., Faithfull, J., Lloyd, N.S., Ellam, R.M., Higgins, J.A. (2018) High-precision 41K/39K measurements by MC-ICPMS indicate terrestrial variability of δ41K. J. Anal. At. Spectrom., 33:175-186.

484 485

36. Maathuis, F.J.M. (2009) Physiological functions of mineral macronutrients. Current Opinion Plant Bio., 12:250-258.

486 487 488

37. Richter, F.M., Mendybaev, R.A., Christensen, J.N., Ebel, D., Gaffney, A. (2011) Laboratory Experiments Bearing on the Origin and Evolution of Olivine-rich Chondrules. Meteoritics and Planet. Sci., 46:1152-1178.

489 490 491

38. Bourg, I.C., Richter, F.M., Christensen, J.N., Sposito, G. (2010) Isotopic MassDependence of Metal Cation Diffusion Coefficients in Liquid Water, Geochim. Cosmochim. Acta, 74:2249-2256.

492 493 494

39. Wang, K. and Jacobsen, S.B. (2016) An estimate of the Bulk Silicate Earth potassium isotopic composition based on MC-ICPMS measurements of basalts. Geochim. Cosmochim. Acta, 178:223-232.

495 496 497

40. Horwitz E. P., McAlister D. R., Bond A. H. and Barrans R. E. (2005) Novel extraction of chromatographic resins based on tetraalkyldiglycolamides: characterization and potential applications. Solvent Extr. Ion Exch., 23:319–344.

498 499 500 501

41. Nielsen, L. C., Druhan, J. L., Yang, W., Brown, S. T., & DePaolo, D. J. (2011). Calcium Isotopes as Tracers of Biogeochemical Processes. In M. Baskaran (Ed.), Handbook of Environmental Isotope Geochemistry (pp. 105–124). Berlin, Heidelberg: Springer Berlin Heidelberg. http://doi.org/10.1007/978-3-642-10637-8_7

502 503

42. Simon, J.I., DePaolo, D.J. (2010) Stable calcium isotopic composition of meteorites and rocky planets. Earth Planet. Sci. Lttrs., 289:457-466.

504 505

43. Jeschke, W.D., Atkins, C.A., Pate, J.S. (1985) Ion circulation via phloem and xylem between root and shoot of nodulated white lupin. J. Plant. Physil., 117:319-330.

506 507

44. Aikio, S., Rämö, K., Manninen, S. (2009) Dynamics of biomass partitioning in two competing meadow plant species. Plant Ecol., 205:129-137.

508 509 510

45. Wiegand, B.A., Chadwick, O.A., Vitousek, P.M., Wooden, J.L. (2005) Ca cycling and isotopic fluxes in forested ecosystems in Hawaii. Geohpys. Res. Lttrs., 32:L11404, doi:10.1029/2005GL022746

511 512 513 514

Wei, R., Guo, Q., Wen, H., Liu, C., Yang, J., Peters, M., Hu, J., Zhu, G., Zhang, H., Tian, L., Han, X., Ma, J., Zhu, C., Wan, Y. (2016) Fractionation of stable cadmium isotopes in the cadmium tolerant Rincinus communis and hyperaccumulator Solanum nigrum. Scientific Reports, 6:24309, doi:10.1038/srep24309

515 516 517 518

47. Peuke, A.D., Rokitta, M., Zimmermann, U., Schreiber, L., Haase, A. (2001) Simultaneous measurement of water flow velocity and solute transport in xylem and phloem of adult plants of Ricinus communis over a daily time course by nuclear magnetic resonance spectrometry. Plant Cell Environ., 24:491-503.

519 520

48. Lobo, V.M.M., Ribeiro, A.C.F., Verissimo, L.M.P. (1998) Diffusion coefficients in aqueous solutions of potassium chloride at high and low concentrations. J. Molecular

ACS Paragon Plus Environment

16

Page 17 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

521

Liquids. 78:139-149.

522 523 524

49. Moynier, F., Fujii, T. (2017) Calcium isotope fractionation between aqueous compounds relevant to low-temperature geochemistry, biology and medicine. Scientific Reports (Nature), 7:44255 doi:10.1038/srep44255

525 526

50. Passioura, J.B. (2010) Plant-water relations. In Encyclopedia of Life Sciences.

527 528 529

51. Doyle, D.A., Cabral, J.M., Pfuetzner, R.A., Kuo, A., Gulbis, J.M., Cohen, S.L., Chait, B.T., Mackinnon, R. (1998) The structure of the potassium channel: Molecular basis of K+ conduction and selectivity. Science, 280:69-77.

530 531 532

52. Lockless, S.W., Zhou, M., MacKinnon R. (2007) Structural and thermodynamic properties of selective ion binding in a K+ channel. PLoS Biol., 5(5):1079-1088. doi:10.1371/journal.pbio.0050121

533 534 535

53. Burger, A., Lichtscheidl, I. (2018) Stable and radioactive cesium: A review about distribution in the environment, uptake and translocation in plants, plant reactions and plants’ potential for bioremediation. Sci. Total Environ., 618:1459-1485.

536 537

54. Bolz, L.H., Mauer, F.A. (1962) Measurement of the lattice constants of neon isotopes in the temperature range of 4-24˚K. Adv. X-ray Anal., 6:242-249.

538 539

55. Hofmann, A.E., Bourg, I.C., DePaolo, D.J. (2012) Ion desolvation as a mechanishm for kinetic fractionation in aqueous systems. Proc. Natl. Acad. Sci., 109:18689-18694.

540 541

56. MacKinnon, R. (2003) Potassium channels. FENS Lettrs., 555:62-65. Doi:10.1016/S0014-5793(03)01104-9.

542 543 544

57. John, S.G., Geis, R.W., Saito, M.A., Boyle, E.A. (2007) Zinc isotope fractionation during high-affinity and low affinity zinc transport by the marine diatom Thalassiosira oceanica. Limnol. Oceanogr., 52(6):2710-2714.

545 546 547 548

58. Deng, T.-H.-B., Cloquet, C., Tang, Y.-T., Sterckeman, T., Echevarria, G., Estrade, N., Morel, J.-L., Qiu, R.-L. (2014) Nickel and zinc isotope fractionation in hyperaccumulating and nonaccumulating plants. Environ. Sci. Technol., 48:1192611933.

549 550

59. Pettigrew, W.T. (2008) Potassium influences on yield and quality production for maize, wheat, soybean and cotton. Pysiologia Plantarum, 133:670-681.

551 552

60. Zörb, C., Senbayram, M., Peiter, E. (2014) Potassium in agriculture- status and perspectives, J. Plant Physiol., 171:656-669.

553 554 555

61. Anschütz, U., Becker, D., Shabala, S. (2014) Going beyond nutrition: Regulation of potassium homoeostatsis as a common denominator of plant adaptive responses to environment. J. Plant Physiol., 171:670-687.

556 557 558 559

62. Santiago Ramos, D.P., Morgan, L.E., Lloyd, N.S., Higgins, J.A. (2018) Reverse weathering in marine sediments and the geochemical cycle of potassium in seawater: Insights from the K isotopic composition (41K/39K) of deep-sea pore-fluids. Geochim. Cosmochim. Acta (In Press) https://doi.org/10.1016/j.gca.2018.02.035

Wiley, Chichester. DOI: 10.1002/9780470015902.a0001288.pub2

ACS Paragon Plus Environment

17

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 29

560 561

63. Wilks, J. (1967) The properties of liquid and solid helium. Clarendon Press, Oxford, 703pp.

562 563 564

64. Buschert, R.C., Merlini, A.E., Pace, S., Rodriguez, S., Grimsditch, M.H. (1988) Effect of isotope concentration on the lattice parameter of germanium perfect crystals. Phys. Rev. B, 38(8):5219-5221.

ACS Paragon Plus Environment

18

Page 19 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

565 566 567 568 569 570 571

ACS Earth and Space Chemistry

Figure 1. The K and Ca isotopic compositions of plants grown hydroponically in this study, and harvested after 2 weeks, 4 weeks, 6 weeks and 8 weeks. Uncertainties shown are ±2s. Top- soybean (roots, stems, leaves). Middle- rice (roots, leaves). Bottom- wheat (roots, leaves). The δ41K and δ44Ca of the K and Ca used in the hydroponic solution are both essentially 0.

ACS Paragon Plus Environment

19

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

572 573 574 575 576 577

Page 20 of 29

Figure 2. A simple box model for the transport of K and Ca in plants, were Fi are fluxes of K or Ca, and εi are isotopic fractionation factors, and ∆i are isotopic differences (e.g. between stem and root). See text for explanation and use of the model, as well as the Supplement for details of its derivation.

ACS Paragon Plus Environment

20

Page 21 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

578

579 580 581 582 583

Figure 3. Plot of plant K+ channels estimated uptake selectivity of Na+, Rb+, and Cs+ relative to K+ (51, 53) under potassium replete conditions vs. ionic radius and the percentage difference in ionic radius relative to K+.

ACS Paragon Plus Environment

21

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

584 585 586 587 588 589

Page 22 of 29

Figure 4. Average atomic mass of the element vs. the fractional difference in atomic radius for the heavier isotope. He (63), Ne (54), K as estimated in the text, and Ge (64). The faded symbol for Ge represents an estimated extrapolation to 2 amu of the fractional difference in size.

ACS Paragon Plus Environment

22

Page 23 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

590

591 592

ACS Earth and Space Chemistry

Table 1. Weeks 2 2 2 2 4 4 4 6 6 6 8 8 8

Sample #1 Wheat Root #1 Soy Root #1 Soy Stem #1 Soy Leaf #2 Soy Root #2 Soy Stem #2 Soy Leaf #3 Soy Root #3 Soy Stem #3 Soy Leaf #4 Soy Root #4 Soy stem #4 Soy Leaf

δ41K -0.74 -0.79 -1.67 -1.21 -0.78 -1.29 -1.25 -0.4 -0.79 -1.23 -0.64 -1.48 -1.13

δ44Ca ±2s ±2s 0.24 -------- -----0.20 -0.73 0.1 0.19 -1.39 0.1 0.26 -1.01 0.1 0.19 -0.68 0.1 0.19 -1.24 0.1 0.14 -0.64 0.1 0.16 -0.2 0.1 0.28 -1.22 0.1 0.23 -0.91 0.1 0.31 -0.39 0.1 0.31 -0.95 0.1 0.21 -0.74 0.1

2 2 4 4 6 6 8 8

#1 Rice Root #1 Rice Leaf #2 Rice Root #2 Rice Leaf #3 Rice Root #3 Rice Leaf #4 Rice Root #4 Rice Leaf

-0.65 -0.73 -0.37 -0.61 -0.28 -0.74 -0.34 -0.34

0.21 ------- ------0.13 ------- ------0.17 -0.32 0.1 0.17 -0.74 0.1 0.17 -0.78 0.1 0.18 -0.8 0.1 0.22 -0.9 0.1 0.21 -0.9 0.1

2 2 4 4 6 6 8 8

#1 Wheat Root #1 Wheat Leaf #2 Wheat Root #2 Wheat Leaf #3 Wheat Root #3 Wheat Leaf #4 Wheat root #4 Wheat leaf K & Ca nitrate

-0.74 0.24 -------- ------1.57 0.19 -1.13 0.1 -0.84 0.20 -0.87 0.1 -1.15 0.19 -0.62 0.1 -0.53 0.3 -0.86 0.1 -1.19 0.21 -0.85 0.1 ------- ------0.67 0.1 -0.51 0.22 -0.92 0.1 0 0.18 -0.03 0.10

Table 2. Averages of the data from table 1. Soybean root Soybean stem Soybean leaf Rice root Rice leaf Wheat root Wheat leaf

Ave. δ41K -0.65 -1.31 -1.216 -0.4 -0.62 -0.74 -1.11

±2s 0.22 0.4 0.095 0.17 0.18 0.36 0.46

Ave. δ44Ca -0.5 -1.2 -0.83 -0.67 -0.81 -0.82 -0.88

±2s 0.26 0.19 0.18 0.4 0.11 0.24 0.22

593

ACS Paragon Plus Environment

23

ACS Earth and Space Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

594 595

Page 24 of 29

Table 3. Comparison on a per amu basis of the fractionation of K and Ca isotopes by plants in this study. Soy root Soy stem Soy leaf Rice root Rice leaf Wheat root Wheat leaf Ave. of all

K isotopic fractionation per amu

±2s

Ca isotopic fractionation per amu

±2s

-0.33 -0.66 -0.61 -0.20 -0.31 -0.37 -0.56 -0.43

0.11 0.15 0.05 0.09 0.09 0.18 0.23 0.14

-0.13 -0.30 -0.21 -0.17 -0.20 -0.21 -0.22 -0.21

0.07 0.05 0.05 0.10 0.06 0.06 0.06 0.05

K/Ca isotopic fractionation per amu 2.6 2.2 2.9 1.2 1.5 1.8 2.5 2.05

±2s 1.6 0.6 0.7 0.9 0.6 1.0 1.2 0.5

596 597

ACS Paragon Plus Environment

24

Page 25 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

598 599

ACS Earth and Space Chemistry

For TOC Only.

600 601

ACS Paragon Plus Environment

25

leaf

stem

root

leaf

stem

root

leaf

stem

root

leaf

stem

Page 26 of 29

Soybean

δ41K ‰ ( ), δ44Ca, ‰ ( )

-0.2

root

0.0

-0.4 -0.6 -0.8 -1.0 -1.2

-1.4 -1.6

leaf

8 Weeks root

leaf

root

leaf

root

6 Weeks

Rice

δ41K ‰ ( ), δ44Ca, ‰ ( )

-0.2

4 Weeks

leaf

0.0

2 Weeks root

-1.8

-0.4 -0.6 -0.8 -1.0 -1.2

-1.4

-0.2

8 Weeks

root

root

leaf

6 Weeks leaf

4 Weeks leaf

leaf

0.0

2 Weeks root

-1.8

root

-1.6

δ41K ‰ ( ), δ44Ca, ‰ ( )

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Earth and Space Chemistry

Wheat

-0.4 -0.6 -0.8 -1.0 -1.2

-1.4 -1.6 -1.8

ACS Paragon Plus Environment

2 Weeks

4 Weeks

6 Weeks

8 Weeks

Page 27 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39

ACS Earth and Space Chemistry

ACS Paragon Plus Environment

ACS Earth and Space Chemistry

-25% 110

-20% 120

-15%

130

-10%

-5%

0%

5%

Ionic Radius, pm

140

150

160

10%

170

15%

Page 28 of 29

20% 180

Rb+

Cs+

Na+

ACS Paragon Plus Environment

25% 190

K+

Selectivity Relative to K+

1 2 3 100 4 1 5 6 7 8 9 10 11 12 13 14 0.1 15 16 17 18 19 20 21 22 0.01 23 24 25 26 27 28 29 30 31 0.001 32 33 34 35 36 37 38 39 40 0.0001 41

Percentage difference in ionic radius from K+

200

Fractional difference in atomic radius for heavier isotope

Page 29 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25

1

0

10

Atomic Mass ACS EarthAve. and Space Chemistry 20

30

40

50

60

70

80

4He-3He

0.1

0.01 22Ne-20Ne

0.001

0.0001

41K-39K (est.)

(Ge est. 2 amu)

0.00001 74Ge-Ge

ACS Paragon Plus Environment

0.000001

(1 amu)