• Keine Ergebnisse gefunden

Nitrate and ammonium differ in their impact on δ

N/A
N/A
Protected

Academic year: 2022

Aktie "Nitrate and ammonium differ in their impact on δ"

Copied!
40
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

Nitrate and ammonium differ in their impact on δ

13

C of plant metabolites and respired CO

2

from tobacco leaves

Shiva Ghiasi1*, Marco M. Lehmann2, Franz-W. Badeck3, Jaleh Ghashghaie4, Robert Hänsch5,6, Rieke Meinen5, Sebastian Streb7, Meike Hüdig8, Michael E. Ruckle1, Dániel

Á. Carrera1, Rolf T. W. Siegwolf1,2, Nina Buchmann1, Roland A. Werner1

1Department of Environmental Systems Science, ETH Zürich, Zürich, Switzerland, 2Forest dynamics, Swiss Federal Institute for Forest, Snow and Landscape Research WSL, Birmensdorf, Switzerland, 3Council for Agricultural Research and Economics, Research Centre for Genomics and Bioinformatics (CREA-GB), Fiorenzuola d´Arda, Italy, 4Laboratoire d'Ecologie, Systématique et Evolution (ESE), Université Paris-Sud, Orsay, France, 5Institute of Plant Biology, Technische Universität Braunschweig, Braunschweig, Germany,

6Center of Molecular Ecophysiology (CMEP), College of Resources and Environment, Southwest University, Chongqing, P.R. China, 7Department of Biology, ETH Zürich, Zürich, Switzerland, 8Institute of Molecular Physiology and Biotechnology of Plants (IMBIO), University of Bonn, Bonn, Germany

*Corresponding author: Shiva Ghiasi Email: shiva.ghiasi@usys.ethz.ch Phone: +41 78 952 5224

This document is the accepted manuscript version of the following article:

Ghiasi, S., Lehmann, M. M., Badeck, F. W., Ghashghaie, J., Hänsch, R., Meinen, R., … Werner, R. A. (2021). Nitrate and ammonium differ in their impact on δ13C of plant metabolites and respired CO2 from tobacco leaves.

Isotopes in Environmental and Health Studies, 57(1), 11-34.

https://doi.org/10.1080/10256016.2020.1810683

(2)

Abstract

The carbon isotopic composition (δ13C) of foliage is often used as proxy for plant 1

performance. However, the effect of NO3- vs. NH4+ supply on δ13C of leaf metabolites and 2

respired CO2 is largely unknown. We supplied tobacco plants with a gradient of NO3- to 3

NH4+ concentration ratios and determined gas exchange variables, concentrations and δ13C of 4

TCA cycle intermediates, δ13C of dark respired CO2, and activities of key enzymes nitrate 5

reductase, malic enzyme and phosphoenolpyruvate carboxylase.

6

Net assimilation rate, dry biomass and concentrations of organic acids and starch decreased 7

along the gradient. In contrast, respiration rates, concentrations of intercellular CO2, soluble 8

sugars and amino acids increased. As NO3- decreased, activities of all measured enzymes 9

decreased. δ13C of CO2 and organic acids closely co-varied and were more positive under 10

NO3- supply, suggesting organic acids as potential substrates for respiration. Together with 11

estimates of intra-molecular 13C enrichment in malate, we conclude that a change in the 12

anaplerotic reaction of TCA cycle possibly contributes to 13C enrichment in organic acids and 13

respired CO2 under NO3- supply. Thus, the effect of NO3- vs. NH4+ on δ13C is highly relevant, 14

particularly if δ13C of leaf metabolites or respiration is used as proxy for plant performance.

15 16

Key words: ammonium (NH4+), nitrate (NO3-), nitrate reductase (NR), phosphoenolpyruvate 17

carboxylase (PEPC), compound specific carbon isotope analysis (CSIA), stable carbon 18

isotopes.

19 20

(3)

Introduction 21

The strong link between carbon (C) and nitrogen (N) within the plant-soil system has been 22

the subject of numerous studies [1–7]. Studying this interaction and its underlying 23

mechanisms has been critical for understanding the biochemical cycling of these elements, 24

particularly during respiration where C skeletons are provided for N assimilation [8]. Nitrate 25

(NO3-) and ammonium (NH4+) are the most important inorganic N species available to plants 26

[9], both following different uptake and assimilation patterns and resulting in fundamentally 27

different plant metabolism. Yet, the potentially different effects of these two inorganic N 28

species on the C isotopic composition (δ13C) of plant metabolites and dark respired CO2 are 29

largely unknown. This is particularly surprising, since increasing atmospheric CO2

30

concentrations have been found to decrease NO3- assimilation [10]. Therefore, a shift to NH4+

31

as a preferable source of N for plants might be expected in the future [11]. Equally important, 32

δ13C values of bulk leaves and leaf metabolites are widely used to study water use efficiency 33

(WUE), plant physiological processes and plant functional responses to climate. If nitrogen 34

assimilation has been or will be changing over time, the proxy of δ13C for WUE might be 35

unreliable. Thus, a better understanding of how different N species availability affects plant C 36

isotope fractionations is urgently needed.

37

NH4+ is described as a paradoxical nutrient because its redox state requires no reduction 38

process prior to its assimilation (via the GS-GOGAT pathway), but most plants with NH4+

39

supply grow slower and produce less biomass compared to those supplied with NO3- [12,13].

40

Former studies have shown physiological toxicity effects caused by NH4+ in plants [14–16], a 41

decrease in carbon assimilation rates [17], changes in the charge balance caused by reduced 42

uptake of cations, such as K+, Mg2+ and Ca2+, and an increased uptake of anions, such as 43

SO42-, PO43- and Cl- [12,18]. Moreover, NH4+ assimilation generates H+ ions which cause a 44

decrease in pH of the cytosol and eventually might result in growth depression [18].

45

(4)

Therefore, NH4+ assimilation takes place almost entirely in the roots [19], fuelled with C 46

skeletons and energy delivered from the leaves [2]. In contrast to NH4+ supply, about 90% of 47

the NO3- ions taken up by the roots are transported to the leaves for assimilation [20] together 48

with K+ as counter-ion [21]. NO3- assimilation involves reduction of NO3- to NH4+ by nitrate 49

and nitrite reductase (NR, NiR) enzymesprior to assimilation via the GS-GOGAT pathway 50

[22]. This reduction is coupled to photosynthesis [23] and to the tricarboxylic acid (TCA) 51

cycle in the mitochondria, providing the reducing power [24,25]. In addition, the TCA cycle 52

delivers energy and C skeletons for N uptake and assimilation [5,13,26]. N assimilation and 53

organic acid concentrations also follow a diurnal cycle. This has been demonstrated by higher 54

activities of NR and PEPC (i.e. the key enzymes in C and N physiology), higher 55

accumulation of malate, NH4+ and glutamine during the first part of the light period and in 56

contrast, a decrease in concentrations of citrate [26]. Given the different effects of both N 57

species on leaf gas exchange, metabolites and enzyme activities, it is likely to expect strong 58

and different effects of these N species on the C isotopic composition of plant material and 59

respired CO2. 60

Within this biochemical framework, the TCA cycle intermediate malate plays a central role.

61

Malate provides reducing equivalents for NO3- reduction as well as a redox storage in the 62

vacuoles [27,28]. It has also been shown that the TCA cycle is not fully functional during the 63

light period and consists of two loosely connected branches producing citrate and malate 64

[8,28–30]. According to Igamberdiev and Bykova [31], during the light period malate has the 65

potential to solely operate the TCA cycle by generating oxaloacetate (OAA) and pyruvate 66

(supplying both substrates of the TCA cycle) via malate dehydrogenase (MDH) and malic 67

enzyme (ME), respectively. Malate is also a key component in replenishing the loss of C 68

skeletons from the TCA cycle during the amino acid metabolism. This replenishment 69

happens through an anaplerotic reaction via phosphoenolpyruvate carboxylase (PEPC) [4]. In 70

(5)

addition, malate maintains the charge balance [32] and is involved in the re-transport of K+ 71

ions released after NO3- assimilation from shoots to the roots [21]. Together with citrate, 72

malate regulates the transfer of reducing power between cell compartments. Thus, both, 73

citrate and malate are expected to be central to unravel the underlying mechanisms if NO3-

74

vs. NH4+ supply affects the δ13C of foliage, metabolites and respired CO2 differently.

75

δ13C values in C3 plants depend primarily on the δ13C values of atmospheric CO2 and 76

photosynthetic isotope fractionation during CO2 assimilation mainly by RuBisCO [33].

77

Physiological responses such as variation in assimilation rates, mesophyll and stomatal 78

conductance can lead to changes in the ratio of leaf intercellular (Ci or more precisely 79

chloroplast CO2 concentration (Cc)) to atmospheric (Ca) CO2 concentrations (Ci/Ca), which in 80

turn changes δ13C of plant organic matter and subsequently that of respired CO2. 81

Consequently, δ13C values of bulk plant material, metabolites and respired CO2 vary under 82

environmental conditions, which alter the Ci/Ca ratio such as drought and/or temperature [34–

83

38], change over time and differ among species [39–47]. Additional “post-photosynthetic 84

isotope fractionations” caused by reactions during plant metabolism (e.g. during dark 85

respiration) also affect the δ13C of plant material [40,48,49]. Particularly, the anaplerotic 86

reaction of the TCA cycle, moderated via PEPC, results in a 13C enrichment of the introduced 87

COOH group (C-4) of OAA which often is converted to the more abundant metabolite like 88

malate [50,51] and eventually δ13C of respired CO2 [38,52]. However, although several 89

studies showed the influence of N source and content on C isotope discrimination during 90

photosynthesis [53–55], to date, there is no study explicitly investigating the effects of N 91

species supplied (changing from NO3- to NH4+) on the δ13C of plant metabolites and of 92

respired CO2. 93

Therefore, the overall aim of this study was to examine the effects of NO3- vs. NH4+ supply 94

on C isotope fractionation of leaf metabolites and dark respired CO2. We hypothesized a 95

(6)

change in δ13C of plant metabolites as well as respired CO2 in response to changing ratios of 96

NO3- and NH4+. Moreover, we hypothesized that malate and the anaplerotic reaction of the 97

TCA cycle play a central role for the fractionation during respiration of NO3- supplied plants.

98

To test our hypotheses, a series of climate chamber based experiments have been conducted 99

and relevant foliar gas exchange variables, enzymatic reactions, as well as concentrations and 100

δ13C values of selected metabolites and respired CO2 of Nicotiana sylvestris leaves have been 101

measured. Plants grew under a gradient of NO3- to NH4+ concentration ratios.

102

Materials and methods 103

Plant material 104

Tobacco seeds (Nicotiana sylvestris) were stratified for three days in water prior to sowing.

105

The seeds were then sown in organic substrate (substrate 2, Klassmann Deilmann, Germany) 106

and grown for two weeks in a walk-in climate chamber with average day/night temperatures 107

of 22/20 °C and relative humidity (RH) of 50-70%. Plants were exposed to 12 hours daylight, 108

supplemented by fluorescent lamps (Sylvania, Switzerland), with an average photosynthetic 109

photon flux density of approximately 400 µmol m-2 s-1 at the leaf surface. After two weeks, 110

each plant was moved to a 750 mL pot filled with sterile sand (Quarzsand A 0.7-1.2 mm, 111

Carlo Bernasconi AG, Switzerland), and perlite (Foma Sa, Switzerland) at the ratio of 5:1.

112

Five plants were used per treatment (see below), grown in individual pots.

113

Nutrients were supplied to plants with nutrient solutions modified according to Arnozis et al.

114

[56] as follows: 1 mM CaCl2; 1 mM MgSO4; 0.25 mM KH2PO4. Microelements and Fe- 115

EDTA were provided with identical concentrations as in a Hoagland solution. In order to 116

study the effect of different N species, 6 mM N was supplied to plants in different ratios of 117

NO3- to NH4+, provided in the forms of Ca(NO3)2, KNO3, (NH4)2SO4, and NH4NO3, while 118

keeping the total amount of N in the solution constant. In the 100% NH4+ nutrient solution an 119

additional 1 mM K2SO4 was added. Throughout the manuscript, when using the term ‘N 120

(7)

species gradient’, we refer to the different ratios of the two inorganic N species supplied.

121

Along this gradient, the following ratios were used: [%NO3--N: %NH4+-N] 100:0 (NO3-), 122

90:10 (90%), 75:25 (75%), 50:50 (50%), 25:75 (25%), 10:90 (10%), 0:100 (NH4+) shown in 123

Table 1. For simplicity, we will use the term “NO3- supply” when referring to plants fed with 124

100% NO3-, 90% NO3-, 75% NO3- and 50% NO3-, and will use the term “NH4+ supply” when 125

referring to plants fed with 75% NH4+, 90% NH4+ and 100% NH4+; sole NO3- and sole NH4+

126

treatments represent 100% NO3- and 100% NH4+ supply, respectively. Plants were treated 127

with the NO3-/NH4+ solutions (pH≈7) every second day. On the days without treatment, water 128

was provided for unlimited water supply. The pH of the drained soil solution was monitored 129

throughout the study. At harvest time, the pH value was ca. 6.5 in NO3- supplied plants and 130

ca. 5 in NH4+ supplied plants.

131

This setup was repeated three times and all plants were grown in identical conditions from 132

seeding, to ensure sufficient leaf material for prospective analyses. One set of plants was used 133

for leaf gas exchange measurements three hours after start of light period. The same leaves 134

were then used for 13C analysis of dark-respired CO2 and for the analysis of δ13C and 135

concentrations in organic matter and metabolites. The second and third sets were also 136

harvested three hours after the start of light period and used for measurements of respiration 137

and enzymatic analyses, respectively. All measurements and final harvest were performed 138

after 45 days of treatment (thus plants were about 8 weeks old at harvest time).

139

Foliar gas exchange variables 140

Net assimilation rate (An), ratio of intercellular CO2 to atmospheric CO2 concentration (Ci/Ca) 141

and stomatal conductance (gs) were measured with a LI-6400 (LI-COR, USA) on one mature, 142

fully expanded leaf attached to the plant. A standard photosynthesis cuvette with a leaf area of 143

6 cm2 (6400-02B LED Light Source) was used while leaf temperature was set to 20 °C. PAR 144

(Photosynthetically Active Radiation) was set at 400 µmol m-2 s-1 and [CO2] (Ca) was kept at 145

(8)

400 µmol mol-1 with a constant flow rate of 500 µmol s-1. After the plants were in steady state 146

(An and gs stabilized), each gas exchange measurement took 60 seconds with the data 147

recorded every 10 seconds and averaged afterwards. The remaining plant leaves were shock- 148

frozen in liquid N2. The dry weight of the leaves was measured after freeze drying to 149

determine plant foliar biomass. These samples were then ground to a fine powder using a ball 150

mill (MM 200, Retsch, Germany) and stored in a desiccator for further analyses.

151

δ13C values of leaf respired CO2 and respiration rates 152

The C isotopic composition of leaf dark-respired CO2 was measured during the day similar to 153

previous studies [52,57,58]. After the measurement of gas exchange, the respective leaf was 154

cut and immediately transferred into a gas tight Tedlar® bag (Keyka Ventures, USA) 155

according to Barbour et al. [59] and Barthel et al. [60]. To avoid the light enhanced dark 156

respiration (LEDR) period, the Tedlar bag containing the leaf was kept in the dark for 30 min 157

before measurements started [36,61–63]. The sampling bag was flushed several times with 158

synthetic CO2-free air (20% O2 and 80% N2, Pangas, Switzerland) until a CO2 free 159

atmosphere was established, controlled by an infrared gas analyser (LI-820, LI-COR, USA).

160

The bag with the leaf was kept in the dark for approximately one hour and the air was 161

sampled with a gas-tight syringe (BD Plastipak, Switzerland) and transferred into a 12 mL 162

evacuated gas-tight glass vial (“Exetainer”, Labco, England). The δ13C value of dark respired 163

CO213CR) was measured with a modified Gasbench II as described by Zeeman et al. [64]

164

coupled to a DeltaplusXP isotope ratio mass spectrometer (IRMS, ThermoFisher, Germany). In 165

addition, the δ13C value of atmospheric CO2 in the climate chamber was determined as -10.6 166

mUr ± 0.9 (SD). All C isotope measurements in CO2 had a typical standard deviation of less 167

than 0.15 mUr.

168

For measurements of respiration rates, plants were also kept in the dark for 30 mins to avoid 169

LEDR. A fully developed leaf was cut and put into a 445 mL glass jar covered with an opaque 170

(9)

plastic foil. The lid of the jar was equipped with a luer-stopcock fitting to allow gas sampling.

171

During a one-hour period after closing the jar, 4 samples were taken. The first sample was 172

taken right after closing the jar, and the following three samples were taken every 20 minutes.

173

CO2 concentrations were measured with a gas chromatograph (456-GC, Scion Instruments, 174

UK). After CO2 sampling, the leaf was scanned (Lide 20, Canoscan, Canon) and the leaf area 175

determined with the software package Fiji [65]. Leaf respiration rates were derived from the 176

linear increase in CO2 concentrations over time. Respiration rates accounted for leaf area, 177

head space volume and temperature (ideal gas law; cf. [66]).

178

δ13C values of bulk organic matter 179

After freeze-drying and milling, approximately 2 mg of bulk leaf material was weighed in Sn 180

capsules (5 x 9 mm, Saentis, CH) for δ13C analysis of bulk leaf organic material (δ13CB) with 181

a Flash EA 1112 Series elemental analyzer (ThermoFisher, Germany) coupled to a 182

DeltaplusXP IRMS via a ConFlo III as described by Brooks et al. [67] and Werner et al. [68].

183

Sequencing and referencing followed the recommendation by Werner and Brand [69]. The 184

measurement precision (SD) of the quality control standard (tyrosine Tyr-Z1) during the δ13C 185

measurements of all EA samples reported in this study was 0.16 mUr.

186

Organic metabolites for bulk δ13C analysis 187

Extraction of organic metabolites was performed according to Lehmann et al. [38]. Briefly, 188

100 mg freeze-dried plant material were transferred to 2 mL reaction tubes. 1.5 mL 85 °C 189

deionized water was added to prevent any enzymatic activity. The samples were then 190

incubated for 30 min in a water bath at 85 °C. After centrifugation for 2 min at 10000 g, the 191

supernatant and the remainder (pellet) were separated and kept for further purifications. Ion 192

exchange chromatography was used to separate the supernatant containing the water-soluble 193

content into sugars, amino acids, and organic acids fractions. 1 mL of the supernatant was 194

added to a column filled with a cation exchanger (Dowex 50WX8, hydrogen form, 100-200 195

(10)

mesh, Sigma Aldrich, CH), that was directly above a second column filled with an anion 196

exchanger (Dowex 1X8, chloride form, 100-200 mesh, Sigma Aldrich, CH). The neutral sugar 197

fraction was eluted with 30 mL deionized water. The amino acid fraction was trapped on the 198

Dowex 50WX8 and then eluted with 30 mL 3 M ammonia solution, while the organic acid 199

fraction was trapped on the Dowex 1X8 column and afterwards eluted with 35 mL 1 M HCl 200

[38,70]. All metabolic fractions were frozen, freeze dried, re-dissolved in deionized water and 201

stored at -20 °C for further use.

202

Starch was enzymatically isolated from the pellet originating after the water extraction 203

mentioned above, according to standard protocols [70,71]. In brief, the pellet was washed 204

several times using an MCW solution (methanol/chloroform/deionized water, 12/5/3, v/v/v) 205

and deionized water, and subsequently dried overnight. On the second day, the starch in the 206

pellet was re-solubilized in water and gelatinized at 100 °C for 15 min and broken down to 207

soluble sugars at 85 °C for 2 hours using the heat resistant -amylase (EC 3.2.1.1, Sigma- 208

Aldrich, CH). The -amylase was separated from the soluble sugars by centrifugation filters 209

(Vivaspin 500, Sartorius, Germany). For the respective δ13C measurements, aliquots from the 210

starch-derived sugar (δ13CSt), sugar (δ13CS), amino acid (δ13Caa), and organic acid (δ13Coa) 211

fractions were pipetted into Sn capsules, dried in an oven at 60 °C, and analysed with the EA- 212

IRMS coupling as described above. Compound specific isotope analysis (CSIA) of individual 213

organic acids was performed following Lehmann et al. [38] and Rinne et al. [72] and CSIA of 214

individual amino acids was performed following Schierbeek [73] and Vlaardingerbroek et al.

215

[74]. Detailed descriptions of these methods are provided in the supplementary information 216

section.

217

Enzymatic activities 218

The leaf material for enzyme activity assays was frozen and immediately ground with pestle 219

and mortar in liquid nitrogen, then stored at -80 C until analyses. The activities of NR, PEPC 220

(11)

and ME were assayed following Dier et al. [75], Gibon et al. [76] and Tronconi et al. [77], 221

respectively. Detailed descriptions of these methods are provided in the supplementary 222

information section.

223

Concentration of carbohydrates 224

Glucose, fructose, sucrose, total sugars and total starch were extracted from the leaves 225

according to Ruckle et al. [78]. In brief, 10 mg of freeze-dried leaf material was washed two 226

times with 80% (v/v) ethanol and two times with 50% (v/v) ethanol to separate the soluble 227

sugars from starch. Each ethanol extract was incubated at 80 °C for 30 min. The remaining 228

pellet containing starch was incubated at 90 °C in 0.2 M KOH and then neutralized with 1 M 229

acetic acid. The starch in the pellet was digested with -amylase and amyloglucosidase 230

(Sigma Aldrich, CH) to convert it to glucose. Starch-derived glucose, glucose, fructose and 231

sucrose were quantified in triplicates using an enzyme-linked assay based on hexokinase, 232

glucose-6-phosphate dehydrogenase, phosphoglucoisomerase, and invertase as described by 233

Ruckle et al. [78]. Sugar dependent conversion of NADP+ to NADPH (Sigma Aldrich, CH) 234

was quantified using a plate reader (Perkin Elmer Enspire©, USA) by measuring the 235

absorption at 340 nm.

236

Statistics and calculations 237

To study the magnitude of post-photosynthetic C isotope fractionation we calculated the 238

apparent respiratory isotope fractionation according to Ghashghaie et al. [79]:

239

dark= (𝛿13CS - 𝛿13CR) / (1 + 𝛿13CR) 240

where δ13CS andδ13CR represent the isotopic composition of soluble sugars and the isotopic 241

composition of respired CO2, respectively. We used the simplified Farquhar model 242

= a + (b-a) * Ci/Ca

243

Where a is the isotope fractionation during diffusion and b during carboxylation to calculate 244

carbon isotope discrimination [34]. We tested the effect of different N species solutions on 245

(12)

enzyme activities, concentrations and δ13C values with one-way ANOVA using R [80]. The 246

different solutions (NO3- to NH4+) were treated as factors. Tukey-HSD post-hoc tests were 247

used for multiple comparisons at the 0.05 significance level. We used linear regression 248

models to calculate the p and R2 values given in the graphs; asterisks represent different 249

significance levels: * for p<0.05, ** for p<0.01 and *** for p<0.001. All δ13C values were 250

measured on V-PDB scale.

251

Results 252

Foliar gas exchange variables 253

The change in N species supplied led to a shift in gas-exchange behaviour in tobacco plants 254

(Tab. 2). We found a significant effect of the N species gradient on net assimilation rate (An), 255

with approximately 50% lower An under NH4+ compared to NO3- supply (Fig. 1A, p=0.012).

256

The ratio of intercellular to atmospheric CO2 concentrations (Ci/Ca) was higher with 257

increasing NH4+ fraction (Fig. 1B, p=0.0014). Although stomatal conductance (gs) was more 258

variable under NH4+ than under NO3- supply, no significant treatment effect was observed 259

(Fig. 1C, p=0.6). In contrast, respiration rate (Rdark) significantly increased when NH4+ supply 260

increased (Fig. 1D, p<0.01), whereas leaf dry biomass significantly decreased with increasing 261

NH4+ (Fig. 1E, p<0.001).

262 263

Carbon isotopic composition of plant metabolites and respired CO2 264

C isotopic composition of leaf bulk organic matter (δ13CB) stayed relatively constant at about 265

-27.4 mUr from 90% NO3- to 10% NO3-, but tended to be lower in 100% NO3- and 100%

266

NH4+ with -28.3 mUr and -29.1 mUr, respectively (Tab. 2, Fig. 2A). Moreover, we did not 267

find a relationship between δ13CB and N content (data not shown). δ13C of leaf metabolites 268

(e.g. sugars, starch and organic acids) and leaf respired CO2 were clearly affected by changes 269

in the N species ratios (Fig. 2). δ13C values of sugars (δ13CS) increased steadily from -27.3 270

(13)

mUr (100% NO3-) to -25.8 mUr (10% NO3-) but decreased sharply to -28.4 mUr under 100%

271

NH4+ (Fig. 2B). Also, δ13C of starch (δ13Cst) increased from -27.2 mUr (100% NO3-) to -25.9 272

mUr (50% NO3-) before decreasing to -27.8 mUr (100% NH4+). No significant differences 273

were observed for isotopic composition of amino acids (δ13Caa: Fig. 2D, p=0.14). δ13C values 274

of organic acids (δ13Coa) and respired CO213CR) followed similar patterns along the N 275

species gradient from NO3- to NH4+, with almost steadily decreasing δ13C values, which were 276

also less negative compared to those of other metabolic fractions particularly under NO3-

277

supply (Figs. 2E,F). Both δ13Coa and δ13CR showed more negative values under NH4+ supply 278

compared to those under NO3- supply, ranging from -24.2 mUr (δ13Coa in 90% NO3-) to -27.8 279

mUr (δ13Coa in 100% NH4+), and from -24.4 mUr (δ13CR in 90% NO3-) to -27.8 mUr (δ13CR in 280

100% NH4+), respectively.

281 282

Compound-specific analysis of organic acids 283

Malate and citrate were the dominant organic acids in the bulk organic acids extracted for 284

compound specific analysis. δ13C values of malate tended to be more negative under NO3-

285

than under NH4+ supply, gradually increasing from -21.6 mUr to -19 mUr, before decreasing 286

again at 100% NH4+ (Tab. 2, Fig. 3A, p=0.07). In contrast, δ13C values of citrate showed a 287

clear effect of N species, with δ13C values decreasing from -22.9 mUr in 90% NO3- to -26.9 288

mUr in 10% NO3- treatment (Fig. 3B, p=0.017). Moreover, very strong effects were observed 289

in malate concentrations, which decreased with increasing NH4+ from 128 µmol g DW-1 in 290

100% NO3- to about 2 µmol g DW-1 in 25% NO3- (Fig. 3G, p<0.001). Similarly, citrate 291

concentrations decreased from 47 µmol g-1 in 75% NO3- to less than 1 µmol g DW-1 in 10%

292

NO3- (Fig. 3H, p<0.001). It has to be noted that concentrations and δ13C values need to be 293

treated with caution since the concentrations of malate and citrate were at the detection limit 294

of HPLC-IRMS measurements for plants supplied with less than 50% NO3- and should be 295

(14)

interpreted with caution.

296 297

Compound specific analysis (CSIA) of amino acids 298

In this study, we focused on four amino acids including alanine (related to pyruvate), serine 299

(a product of photorespiration in leaves), asparagine/aspartic acid (related to OAA) and 300

glutamine/glutamic acid (related to 2-oxoglutarate). N species treatments significantly 301

influenced δ13C values of alanine, asparagine and serine (Fig. 3C-E, p<0.001 for alanine and 302

serine, p=0.005 for asparagine), but not those of glutamine (Fig. 3F). While δ13C of alanine 303

was gradually decreasing with increasing NH4+ supply, with highest values in 100% NO3- (- 304

27.9 mUr) and lowest in 100% NH4+ (-29.5 mUr, Fig. 3C), no clear pattern was observed 305

along the N species gradient for asparagine (Fig. 3D). Similar to alanine, serine was more 306

enriched in 13C in plants grown under NO3- supply, with highest δ13C value in 90% NO3- (- 307

27.8 mUr), and more depleted in 13C under NH4+ supply, with the lowest δ13C value in 10%

308

NO3- (-32.5 mUr, Fig. 3E). Furthermore, concentrations of all four amino acids increased 309

from NO3- supply towards 10% NO3-, before sharply decreasing again at 100% NH4+, almost 310

reaching the same values as under NO3- supply (Figs. 3 I-L, p<0.01). The N species effect 311

was very clear in alanine and also in serine, with the highest concentrations of alanine in 10%

312

NO3- (2.8 µmol g DW-1) and lowest in 100% NO3- (0.5 µmol g DW-1) and highest 313

concentrations of serine in 10% NO3- (23.7 µmol g DW-1) and lowest in 90% NO3- (2.6 µmol 314

g DW-1). The influence of N species on the concentrations of asparagine and glutamine was 315

not as clear, although NH4+ supplied plants showed higher concentrations compared to NO3-

316

supplied plants.

317 318

Enzyme activity 319

Enzymatic activities of NR, PEPC and ME were negatively affected by the gradient of NO3-

320

(15)

to NH4+ supply (Fig. 4, p<0.001), most clearly seen in NR and PEPC. As the fraction of NH4+

321

increased, NR activity decreased from 1.9 nmol NO2- g FW-1 s-1 (100% NO3-) to 0.7 nmol 322

NO2- g FW-1 s-1 (10% NO3-, Fig. 4A). Similarly, PEPC activity decreased from 9.5 nmol 323

NAD g FW-1 s-1 (100% NO3-) to 4.3 nmol NAD g FW-1 s-1 (100% NH4+, Fig. 4B). In contrast, 324

ME activity was lowest under 25% NO3- (269 nmol NAD g FW-1 s-1) and highest under 50%

325

NO3- supply (505 nmol NAD g FW-1 s-1, Fig. 4C). The activities of NR and ME in plants 326

growing under 100% NH4+ could not be analysed due to insufficient amount of plant 327

material.

328 329

Concentration of carbohydrates 330

The gradient of N species from NO3- to NH4+, strongly influenced concentrations of both 331

water-soluble carbohydrates and starch (Fig. 5). Concentrations of glucose, fructose, sucrose 332

and total sugars significantly increased as the percentage of NH4+ in the N supply increased, 333

except under 100% NH4+ (Fig. 5A-D). In contrast, starch concentrations were highest in 334

plants supplied with higher percentage of NO3-, with concentrations decreasing from 359 mg 335

g DW-1 in 50% NO3- to 95 mg g DW-1 under 10% NO3- supply (Fig. 5E).

336 337

Potential substrates for respiration and apparent isotope fractionation 338

As stated above, leaf respiration rates as well as δ13CR were influenced by the gradient of N 339

species. In addition, we found a significant and positive relationship between δ13CR and 340

δ13Coa, with a slope close to one and an intercept close to zero, representing a 1:1 relationship 341

between these two variables (Fig. 6, R2=0.93, p<0.001, y = 0.97x - 0.72). This relationship 342

clearly showed that both organic acids and respired CO2 became more depleted in 13C along 343

the N species gradient from NO3- to NH4+. 344

Also, apparent isotope fractionation (ℯdark), calculated from δ13CS and δ13CR, was influenced 345

(16)

by the N species gradient, with negative ℯdark values indicating higher 13C enrichment in 346

respired CO2 compared to sugars as the first products of photosynthesis. ℯdark in plants 347

supplied with more than 50% NO3- was more negative compared to plants supplied 348

dominantly with NH4+ (Fig. 7). ℯdark values were the lowest (-2.4 mUr) in plants of the 90%

349

NO3- treatment and increased along the gradient to the highest values in plants under 10%

350

NO3- (0.9 mUr).

351

Discussion 352

NO3- to NH4+ gradient affects foliar gas exchange and metabolism 353

Our main aim was to study C isotope fractionation in response to a supply gradient of two 354

different N species. We therefore setup an experiment with tobacco plants supplied with a N 355

gradient from 100% NO3- to 100% NH4+. As expected, foliar gas exchange characteristics 356

and biomass were strongly influenced by this N supply gradient (Fig. 1). As NO3- decreased, 357

net assimilation rate also decreased in our tobacco plants (Fig. 1A), similar to earlier studies 358

with spinach [81], sugar beet [82] and maize [83]. The decreasing net assimilation rate and 359

the simultaneously increasing respiration rate along the gradient (Fig. 1D), can explain the 360

observed lower biomass in plants under NH4+ supply (Fig. 1E), most likely a sign of NH4+

361

toxicity caused by NH4+ supply [16, 12, 18], supported by the lower pH values (by 1.5) we 362

observed in the drained soil solution from NH4+ supplied plants. Moreover, lower 363

assimilation rate with decrease in NO3- supply is in line with the lower needs for reducing 364

equivalents.

365

We observed strong impacts of the ratio of N species supplied on all measured metabolite 366

concentrations such as organic acids (malate and citrate), soluble sugars, starch and amino 367

acids (alanine, serine, asparagine and glutamine), as well as on measured enzyme activities.

368

As suggested by the higher assimilation of plants dominantly supplied with NO3-, there was 369

probably a higher supply of reducing power, in line with the observed higher NR activity 370

(17)

(Fig. 4A). Decreasing PEPC and ME activities (Fig. 4B,C) together with decreasing 371

concentrations of organic acids along the gradient from NO3- to NH4+ (Fig. 3G,H) imply an 372

important role of organic acids (e.g. malate) and the anaplerotic reaction of the TCA cycle 373

under NO3-, but not under NH4+ supply [5,29,55]. Altogether, these findings clearly 374

confirmed that a change in N species supplied to plants has an influence on carbon 375

metabolism and thus on leaf respiratory metabolism.

376

We further observed higher concentrations of soluble sugars in the leaves of plants supplied 377

dominantly with NH4+ (Fig. 5D) which aligned with higher respiration rates in these plants 378

(Fig. 1D), resulting in a strong positive relationship between sugar concentrations and 379

respiration (R2=0.75, p<0.05, Fig. S1). This might be a stress response to NH4+ toxicity, 380

supported by suppressed growth observed in this study as also suggested by Hachiya et al.

381

[84]. N assimilation under high NH4+ supply takes place mostly in the roots [17,19] and 382

requires C skeletons to be delivered from the leaves [85]. However, not only aboveground 383

but also root biomass was low in NH4+ supplied plants (data is not shown) [86].

384

Moreover, the N species gradient not only influenced concentrations of organic acids and 385

carbohydrates, but also the concentrations of individual amino acids (Fig. 3). We observed an 386

increase in concentrations of alanine, serine, asparagine and glutamine with increasing NH4+

387

in the supply solutions (Fig. 3 I-L), supporting previous findings [2,22,87]. These 388

observations might simply be signs of a detoxification strategy, i.e., remove free NH4+ ions 389

from the cytosol by converting them into amino acids [19,22,88]. Moreover, decreasing C:N 390

ratios are reported to relate to increasing amino acid biosynthesis [24,87], as seen for the 391

NH4+ supplied plants in our study (Fig. S2 A). Consequently, we are confident that our 392

experimental setup allowed in-depth assessment of the C isotopic composition of metabolites 393

and respired CO2 in relation to a N species gradient.

394 395

(18)

NO3- to NH4+ gradient influences δ13C of amino acids 396

Moreover, we conducted compound specific isotope analysis of individual amino acids to 397

investigate if changes observed in the post-photosynthetic fractionation between NO3- and 398

NH4+ supplied plants were mirrored in the δ13C of individual amino acids as the products of 399

N assimilation. Analysis of amino acids showed different effects of the two N species on δ13C 400

values of individual amino acids, particularly of alanine and serine (Fig. 3C,E; except for 401

serine under 100% NH4+ supply). A possible explanation for 13C enriched alanine in NO3-

402

supplied plants can be the isotope fractionation caused by pyruvate dehydrogenase (PDH), 403

resulting in an accumulation of 13C enriched glycolytic pyruvate that is known as a precursor 404

for alanine biosynthesis [89,90]. Alternatively, the observed 13C enriched alanine in NO3-

405

supplied plants can be explained by higher ME activity (Fig. 4C), supplying amino acid 406

biosynthesis with 13C enriched pyruvate from anaplerotically synthesized malate.

407

δ13C values of serine and alanine were similar in NO3- supplied plants (around -27 mUr), but 408

serine was by 3 mUr more 13C depleted in NH4+ plants compared to alanine (-32.5 mUr to - 409

29.5 mUr, respectively). Tcherkez [91] suggested a 13C depletion in serine molecules 410

compared to glycine in relation to photorespiration. Our findings therefore can convey 411

possible higher photorespiration in NH4+ supplied plants, resulting in 13C depletion of serine.

412

Although δ13C of asparagine was not influenced by the gradient of N species, it showed the 413

highest 13C enrichment among all other amino acids. Sechley et al. [92] underlined the role of 414

OAA/malate as donors of C skeleton for synthesis of asparagine. Given that malate was the 415

most 13C enriched compound in our study (Fig. 3A), a contribution of this compound to the 416

13C enrichment of asparagine is likely.

417 418

NO3- to NH4+ gradient influences carbon isotope fractionation 419

(19)

We observed distinct changes in δ13C values of respired CO2 and various metabolites such as 420

soluble carbohydrates, starch, organic acids and consequently leaf bulk organic matter in 421

relation to the N species gradient (Fig. 2). Despite the clear increase of Ci/Ca ratio from NO3-

422

to NH4+ supplied plants (Fig. 1B), suggesting larger 13C discrimination and therefore more 423

13C depleted assimilated along the gradient, the δ13C values of sugars, starch and bulk organic 424

matter showed no significant response (Fig. 2A,B). Therefore, we could not explain the 425

differences between δ13C values of metabolites (sugars and starch as the first products of 426

photosynthesis) along the gradient (R2<0.54) using the Ci/Ca ratio alone, based on the 427

simplified Farquhar model (= a + (b-a) * Ci/Ca, Fig. S3, [34]). This might have several 428

reasons: (1) N concentration positively influences mesophyll conductance and thus Cc/Ca, 429

resulting in higher 13C discrimination which in turn causes more 13C depleted sugars 430

[43,45,93,94]. However, although we found higher N concentration in the leaves of NH4+

431

supplied plants (as shown in Fig. S2 B) and thus potentially increased mesophyll 432

conductance, soluble sugars in NH4+ supplied plants were not depleted in 13C. (2) The foliar 433

sugar pool has accumulated over longer time periods than can be represented by individual 434

snap-shot Ci/Ca measurements. (3) According to Caemmerer and Evans [43], there is a 435

possible influence of PEPC activity on the expected C isotope discrimination during 436

carboxylation by RuBisCo. Considering the different PEPC activities measured in NO3- and 437

NH4+ supplied plants, such an effect of PEPC on discrimination cannot be excluded.

438

To further assess the mechanisms underlying the effects of NO3- vs. NH4+ supply on the 439

fractionation during respiration, we calculated the apparent isotope fractionation (ℯdark) 440

between the sugars δ13CS (as the first photosynthetic product) and respired CO213CR) 441

according to Ghashghaie et al. [79]. Interestingly, ℯdark values were much more negative in 442

plants supplied with more than 50% NO3- than in plants supplied predominantly with NH4+

443

(Fig. 7). This implies that in plants supplied with NO3-, substrates for the respiration were 444

(20)

more enriched in 13C relative to sugars (as the first products of photosynthesis and main 445

metabolites in the respiratory pathway) which in turn resulted in the observed 13C enrichment 446

of respired CO2. In contrast, plants with higher percentage of NH4+ supply showed ℯdark

447

values which were on average close to or slightly above zero, reflecting a very minor and 448

thus negligible difference between δ13CS and δ13CR (Fig. 7). These results confirmed our 449

hypothesis that a shift from NO3- to NH4+ supply mainly influences post-photosynthetic 450

isotope fractionation.

451

The high concentrations of malate and the high PEPC activity in NO3- supplied plants 452

together with relatively high δ13C values of malate amongst all other metabolites (Fig. 3A) 453

further support the idea that the anaplerotic reaction of the TCA cycle could play an 454

important role in the 13C enrichment of CO2 respired by the leaves of NO3- supplied plants. In 455

order to determine the relative contribution of the anaplerotic reaction responsible for the 13C 456

enrichment in leaf-respired CO2, we calculated the possible 13C enrichment of the C-4 atom 457

in malate [95]. According to the model proposed by Mezler and O’Leary [95], the carbon 458

introduced into malate via PEPC is about 6 mUr more enriched in 13C than that of 459

intercellular (internal) CO2. After calculating the δ13C of intercellular CO2, we adjusted the 460

Melzer and O’Leary model by assuming that in plants supplied with 100% NO3-, PEPC has 461

the highest activity (supported by our own results, Fig. 4) and therefore these 100% NO3-

462

plants have the highest enrichment of the C-4 atom in malate (i.e. an additional 6 mUr). The 463

13C enrichment of all other N treatments along the gradient were then calculated by using the 464

ratio of their PEPC activity compared to that of 100% NO3- supplied plants (Fig. S4). Our 465

results showed a clear decrease in the 13C enrichment of the C-4 atom in malate along the N 466

species gradient. These findings indicated that a higher proportion of anaplerotically 467

introduced carboxyl group(s) in malate might be released as CO2 during amino acid 468

biosynthesis or plant dark respiration, particularly in NO3- supplied plants (Fig. 8). We 469

(21)

therefore conclude that such a change in the PEPC related mechanism along the N species 470

gradient is the most likely pathway to explain 13C enrichments observed in δ13C of organic 471

acids and of respired CO2 under NO3- supply (see Fig. 6).

472 473

Conclusion 474

This study aimed to understand the influence of a supply gradient in NO3- and NH4+ supply 475

on the C isotope fractionation as seen in leaf metabolites and during leaf respiration. Several 476

lines of evidence suggest that modifications of the N species supplied to plants, which 477

changed foliar gas exchange, enzymatic activities and concentrations of metabolites (i.e.

478

sugars, amino acids and organic acids), also changed the C isotopic composition of leaf 479

metabolites and respired CO2, supporting our first hypothesis. With our second hypothesis, 480

we suggested a major role of the anaplerotic reaction of the TCA cycle (via PEPC) and 481

organic acids. Indeed, our results indicated a higher contribution of anaplerotically 482

assimilated carbon in the process of dark respiration in plants supplied dominantly with NO3-, 483

leading to enrichments in metabolites and respiratory CO2. Our findings on the impact of N 484

species on δ13C of organic acids and respired CO2 might gain additional relevance since the 485

N species available to vegetation is strongly affected by either management or anthropogenic 486

pollution while δ13C values of leaf metabolites or respiration are widely used in plant 487

sciences, e.g. as a proxies for plant physiological responses to environmental change or to 488

partition ecosystem CO2 fluxes.

489 490

Acknowledgments 491

The authors thank A. H. Khoshgoftarmanesh from Isfahan University of Technology for 492

helping with the calculation of the salts needed for the gradient of nutrient solution. We also 493

acknowledge the help of A. Ackermann for EA-IRMS measurements in the isotope 494

(22)

laboratory of the Grassland Sciences Group at ETH Zurich. We appreciate statistical help 495

given by C. Bachofen from the Grassland Sciences Group. We would like to thank B. Studer 496

for use of laboratory space, reagents, and equipment for the carbohydrate analysis from the 497

Molecular Plant Breeding group at ETH Zurich. We would like to thank H.J Lendi (ETH 498

Zurich) for technical support of the climate chambers and freeze-dryers.

499 500

Declaration of authorship 501

SG, RAW, NB, FWB, JG, MML, RTWS developed the presented idea, planned, and designed the 502

research. SG conducted the experiments as well as data analysis and wrote the manuscript. SG, RH, 503

RM, SS, MH performed enzyme activity assays and MER and DAC contributed in carbohydrate 504

concentration measurements. MML carried out compound specific analysis of organic acids. All 505

authors discussed the results and contributed to the final manuscript.

506 507

Funding 508

SG was funded by the Swiss National Science Foundation (project 205321_153545 “CarIN”) granted 509

to RAW and RTWS. This research was partially supported by the COOP—Research Fellowship 510

Program (HERC) through the ETH-World Food System Center and the Swiss Federal Office for 511

Agriculture (FOAG) granted to MER and by the SNF Ambizione project “TreeCarbo”

512

(PZ00P2_179978) granted to MML.

513 514

Supplemental Material 515

The following supplemental materials are available.

516

Figure S1. Linear regression between respiration rate (µmol m-2 s-1) and concentration of 517

soluble sugars (mg g DW-1) 518

Figure S2. Carbon to nitrogen ratio (C:N) and nitrogen concentration (%) 519

(23)

Figure S3. Linear regression between measured 𝛿13C of starch/sugar and modeled 𝛿13C 520

values using the simplified Farquhar model 521

Figure S4. Modelled carbon isotopic composition of C-4 of malate in the leaves of tobacco 522

plants 523

524

Methods of compound-specific δ13C and concentration analysis of amino acids and organic 525

acids 526

Enzyme activity analysis: NR, PEPC and ME 527

528

Conflict of Interest Statement 529

The authors declare that the research was conducted in the absence of any commercial or 530

financial relationships that could be construed as a potential conflict of interest.

531

532

(24)

References

[1] Berveiller D, Fresneau C, Damesin C. Effect of soil nitrogen supply on carbon assimilation by tree stems. Ann. For. Sci. 2010;67:609–609.

[2] Cramer MD, Lewis O a M. The influence of NO3- and NH4+ nutrition on the carbon and nitrogen partitioning characteristics of wheat (Triticum aestivum L.) and maize (Zea mays L.) plants. Plant Soil. 1993;154:289–300.

[3] Flores P, Hellín P, Fenoll J, et al. Carbon isotopic discrimination in pepper seedlings as affected by nitrogen level and source. Acta Hortic. 2012;938:273–276.

[4] Huppe HC, Turpin DH. Integration of carbon and nitrogen metabolism in plant and algal cells.

Annu. Rev. Plant Physiol. Plant Mol. Biol. 1994;45:577–507.

[5] Nunes-Nesi A, Fernie AR, Stitt M. Metabolic and signaling aspects underpinning the regulation of plant carbon nitrogen interactions. Mol. Plant. 2010;3:973–996.

[6] Poorter H, Remkes C, Lambers H. Carbon and nitrogen economy of 24 wild species differing in relative growth rate. Plant Physiol. 1990;94:621–627.

[7] Raab TK, Terry N. Carbon, nitrogen, and nutrient interactions in Beta vulgaris L. as influenced by nitrogen source, NO3-, versus NH4+. Plant Physiol. 1995;107:575–584.

[8] Tcherkez G, Mahe A, Gauthier P, et al. In folio respiratory fluxomics revealed by 13C isotopic labeling and H/D isotope effects highlight the noncyclic nature of the tricarboxylic acid cycle in illuminated leaves. Plant Physiol. 2009;151:620–630.

[9] Marschner P. Marschner’s mineral nutrition of higher plants. 3rd ed. Marschner’s Miner. Nutr.

High. Plants. Academic Press London; 2012.

[10] Bloom AJ, Burger M, Kimball BA, et al. Nitrate assimilation is inhibited by elevated CO2 in field-grown wheat. Nat. Clim. Chang. 2014;4:477–480.

[11] Hachiya T, Sakakibara H. Interactions between nitrate and ammonium in their uptake, allocation, assimilation, and signaling in plants. J. Exp. Bot. 2016;68:1–12.

[12] Britto DT, Kronzucker HJ. NH4+ toxicity in higher plants: a critical review. J. Plant Physiol.

2002;159:567–584.

[13] Setién I, Vega-Mas I, Celestino N, et al. Root phosphoenolpyruvate carboxylase and NAD- malic enzymes activity increase the ammonium-assimilating capacity in tomato. J. Plant Physiol. 2014;171:49–63.

[14] Britto DT, Siddiqi MY, Glass ADM, et al. Futile transmembrane NH4+ cycling: A cellular hypothesis to explain ammonium toxicity in plants. PNAS. 2001;98:4255–4258.

[15] Esteban R, Ariz I, Cruz C, et al. Review: Mechanisms of ammonium toxicity and the quest for tolerance. Plant Sci. 2016;248:92–101.

[16] Walch-Liu P, Nter Neumann G, Bangerth F, et al. Rapid effects of nitrogen form on leaf morphogenesis in tobacco. J. Exp. Bot. 2000;51:227–237.

[17] Setién I, Fuertes-Mendizabal T, González A, et al. High irradiance improves ammonium tolerance in wheat plants by increasing N assimilation. J. Plant Physiol. 2013;170:758–771.

[18] Andrews M, Raven JA, Lea PJ. Do plants need nitrate? The mechanisms by which nitrogen form affects plants. Ann. Appl. Biol. 2013;163:174–199.

[19] Marschner H. Mineral nutrition of higher plants. 2nd ed. Academic Press New York; 1995.

[20] Hänsch R, Fessel DG, Witt C, et al. Tobacco plants that lack expression of functional nitrate reductase in roots show changes in growth rates and metabolite accumulation. J. Exp. Bot.

2001;52:1251–1258.

[21] Ben-Zioni A, Vaadia Y, Lips SH. Nitrate uptake by roots as regulated by nitrate reduction products of the shoot. Physiol. Plant. 1971;24:288–290.

[22] Kandlbinder A, Da Cruz C, Kaiser WM. Response of primary plant metabolism to the N-

Referenzen

ÄHNLICHE DOKUMENTE

[124] reported that tetracyclines (oxytetracycline (OTC) and CTC) were the most common antibiotics present in swine, beef, and turkey manures. Sulfonamides are among the most

In broad terms, however, the US remains a major and viable world power, and the emergence of other states does not mean the US role in the Middle East and the Gulf is

The AP diameters of the caudal endplate of the upper vertebral body and the cranial endplate of the lower vertebral body were used to determine whether disproportion ex- isted

If the disease has attacked and destroyed the structure of papillary lines in the epidermis layer of the skin, the papillary lines will not grow in the same form as before (if at

Its focus of research concerns the logical interconnections and relationships which combine the Informatics, Information Technology (IT) and Information Systems (IS) fields with

6 Context-free grammars are systems of Reg-equations 22 7 Extending the derivative parser to parsers for CFGs 24.. Constructor and destructor signatures, Reg

Abstract: I analyze company news from Reuters with the ‘General In- quirer’ and relate measures of positive sentiment, negative sentiment and disagreement to abnormal stock

Table 1 shows the problems arising in practice by transition from an improvement policy to basic technological changes in market production, research and development, and