• Nie Znaleziono Wyników

Coupling Stable Isotope Analysis with Gas Push-Pull Tests to Derive In Situ Values for the Fractionation Factor αox Associated with the Microbial Oxidation of Methane in Soils

N/A
N/A
Protected

Academic year: 2021

Share "Coupling Stable Isotope Analysis with Gas Push-Pull Tests to Derive In Situ Values for the Fractionation Factor αox Associated with the Microbial Oxidation of Methane in Soils"

Copied!
31
0
0

Pełen tekst

(1)

Delft University of Technology

Coupling Stable Isotope Analysis with Gas Push-Pull Tests to Derive In Situ Values for the Fractionation Factor αox Associated with the Microbial Oxidation of Methane in Soils

Gebert, Julia; Streese-Kleeberg, J. DOI

10.2136/sssaj2016.11.0387 Publication date

2017

Document Version

Accepted author manuscript Published in

Soil Science Society of America. Journal

Citation (APA)

Gebert, J., & Streese-Kleeberg, J. (2017). Coupling Stable Isotope Analysis with Gas Push-Pull Tests to Derive In Situ Values for the Fractionation Factor αox Associated with the Microbial Oxidation of Methane in Soils. Soil Science Society of America. Journal, 81(5), 1107-1114.

https://doi.org/10.2136/sssaj2016.11.0387 Important note

To cite this publication, please use the final published version (if applicable). Please check the document version above.

Copyright

Other than for strictly personal use, it is not permitted to download, forward or distribute the text or part of it, without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license such as Creative Commons. Takedown policy

Please contact us and provide details if you believe this document breaches copyrights. We will remove access to the work immediately and investigate your claim.

(2)

Coupling stable isotope analysis with gas push-pull tests

1

to derive in-situ values for the fractionation factor

α

ox

2

associated with the microbial oxidation of methane in

3

soils

4

KEYWORDS 5

Landfill cover soil, greenhouse gas emission, methane oxidation, isotopic fractionation 6

ABSTRACT 7

Prerequisite for the application of stable isotope fractionation for the quantification of the 8

methane oxidation efficiency of landfill covers is that the fractionation factor αox is known or can

9

be estimated with adequate accuracy. So far, αox has only been determined in laboratory

10

experiments. For the first time, αox was determined under in situ conditions in the field by

11

coupling two independent methods, gas push-pull tests and stable isotope analysis, to assess 12

biological fractionation of methane isotopologues in landfill cover soils. On six landfills with 13

nine points of investigation, 22 measurements were carried out, covering a wide range of 14

environmental conditions such as soil temperature and moisture and observed oxidation rates. 15

Values for αox varied between near 1, indicating only little fractionation, and 1.151. Correlation

16

of αox with the methane oxidation rate found by gas push-pull tests revealed a clear asymptotic

(3)

relationship with low rates being associated with high values for αox and high rates resulting in

18

only little fractionation. Values for αox varied between the different landfills, but also between

19

the individual points of investigation on the same landfill. The latter is assumed to reflect the 20

spatial variability of methanotrophic activity due to spatial variability in soil moisture and hence 21

air-filled porosity, as well as the spatial variability of gas fluxes. Further, significant variation of 22

αox was observed also for the same sampling point, presumably reflecting the temporal

23

variability of factors influencing methanotrophic activity. These effects could include seasonally 24

changing environmental conditions such a soil temperature and moisture, but also the temporal 25

variability of gas fluxes through the landfill soil cover, changing exposure of methanotrophs to 26

methane and oxygen and hence their activity. The quantification of the methane oxidation 27

efficiency using fractionation of stable isotopes is very sensitive to αox. Assuming a value

28

constant in time and space and transferring this value from laboratory experiments to field 29

settings entails significant uncertainty regarding the quantification of methane oxidation. 30

(4)

1 INTRODUCTION

31

Landfills are estimated to be the second to third largest source of anthropogenic methane 32

emissions (EEA, 2009; USEPA, 2016). Sanitary landfilling including active extraction and 33

subsequent treatment or energetic utilization of landfill gas is mandatory in many countries. The 34

microbial oxidation of methane in optimized landfill covers, windows and filters is an option to 35

complement technical measures and mitigate residual methane fluxes for which technical 36

treatment is no longer feasible (Bogner et al., 2007; Huber-Humer et al., 2008; Scheutz et al., 37

2009). It is also an option for the many old landfills that are not equipped with a gas extraction 38

system or for landfills where gas production is low from the start. The potential of biological 39

treatment is widely recognized. Both operators and regulators, however, require accepted 40

methods to prove and quantify the mitigation effect. 41

The fractionation of stable isotopes (SI), naturally occurring as a result of the oxidation process, 42

offers one of the few approaches to assess the efficiency of microbial methane oxidation (fox) in

43

situ (Chanton et al., 1999, 2011; Börjesson et al., 2001, 2007; Cabral et al., 2010; Abichou et al., 44

2011; Widory et al., 2012). The method is based on the fact that carbon naturally exists in the 45

form of two stable isotopes (12C and 13C) whose size, steric properties and diffusion coefficient 46

differ, leading to unequal behavior at enzyme binding sites (Jahnke et al., 1999) and to different 47

transport rates. Due to the preferential oxidation (Barker and Fritz, 1981) and faster diffusive 48

transport (Tyler et al., 1994; Reeburgh et al., 1997; De Visscher et al., 2004) of the lighter 49

isotope, the heavier isotope is enriched in the remaining gas phase. The factors by which 50

oxidation and transport fractionate, αox and αtrans, respectively, are used to calculate methane

51

oxidation efficiency from the difference in isotopic signature between the landfill gas and the gas 52

emitted from the soil surface or any other point of interest. In the past, values of αox have been

(5)

determined in laboratory under standardized conditions and have then been used to calculate 54

methane oxidation efficiencies from field isotopic signature data. However, as αox is known to

55

vary with oxidation rate (Templeton et al., 2006; Chanton et al., 2008), temperature (Chanton et 56

al., 2008), and possibly the composition of the methanotrophic community (Jahnke et al., 1999; 57

Templeton et al., 2006), the value was hypothesized to be subject to site-specific and seasonal 58

variation. 59

In order to determine αox in situ, the oxidation efficiency must also be determined in situ. This is

60

possible using the gas push-pull test (GPPT) methodology (Gómez et al., 2009; Urmann et al., 61

2009; Streese-Kleeberg et al., 2011). During a GPPT, a defined volume of the reactive gas of 62

interest (in this case methane and oxygen) and a conservative tracer with similar transport 63

properties (e.g. argon) are injected into the soil’s vadose zone. Following injection, the mixture 64

of reactant, tracer and soil gas is extracted via the same tube. During extraction, the gas mixture 65

is sampled periodically in order to obtain breakthrough curves of reactant and tracer. The 66

reaction rate is calculated from the difference in the breakthrough curve of the reactive gas 67

compared to the one of the tracer. The combination of GPPTs with SI analyses allows the in situ 68

determination of αox by simultaneously collecting data on the change of the isotopic composition

69

and on the oxidation efficiency. The aim of the study was to verify the range of fractionation 70

factors previously published from laboratory experiments and to investigate the hypothesized in-71

situ variation of the fractionation factor resulting from to the variation in field oxidation rates. 72

(6)

2 MATHEMATICAL BACKGROUND OF THE STABLE ISOTOPE APPROACH

73

The shift in the isotopic ratio of 13C to 12C through microbial methane oxidation or through 74

diffusive gas transport, signified in the delta notation as δ13C, is described by the following 75

equation: 76

δ13

C [‰] = ((Rsample/Rstandard) – 1)*1000 Eq. 1

77

Where Rsample= 13C/12C ratio in the sample

78

Rstandard= 13C/12C ratio of the carbon reference standard VPDB (Vienna Peedee

79

Belemnite, 13C/12C = 0.0112372). 80

The fractionation factors effective during the oxidation process and during gas transport are 81

termed αoxand αtrans, respectively, and are defined as the oxidation or transport rate constant for

82

12

CH4 divided by the oxidation or transport rate of 13CH4:

83

αox, trans = k12/k13 Eq. 2

84

No fractionation is assumed when gas transport is advective, then αtrans is equal to 1.

85

The exact factors for oxidative and transport fractionation need to be known to quantify methane 86

oxidation efficiency from the measured shift in stable isotope ratios as follows (Blair et al., 87

1985): 88

𝑓𝑓𝑜𝑜𝑜𝑜 =1000 × (𝛼𝛼(𝛿𝛿𝑠𝑠−𝛿𝛿𝑎𝑎𝑎𝑎𝑎𝑎𝑎𝑎𝑎𝑎𝑎𝑎−𝛼𝛼𝑡𝑡𝑡𝑡𝑎𝑎𝑎𝑎𝑠𝑠) ) Eq. 3

(7)

Where fox = fraction of methane load oxidized at the particular sampling location, also

90

defined as methane oxidation efficiency 91

δs = δ13C value in sample

92

δanox = δ13C value at the chosen point of reference, often the anaerobic zone of

93

methane production 94

αox, αtrans = factors for oxidative and diffusive fractionation.

95

As a fraction of 1 or a percentage (if multiplied by 100), fox represents a relative value, not an

96

absolute oxidation rate. 97

3 MATERIALS AND METHODS

98

2.1 Site description

99

Investigations were carried out on five old non-sanitary municipal solid waste landfills (landfills 100

A, D, L, H and K) in north western Germany, all filled in the 1960s to the 1980s. The landfills 101

were part of a baseline study on methane oxidation in non-optimized cover soils, conducted 102

within the framework of the MiMethox project (Rachor et al., 2009). The landfills were covered 103

with whatever soil was available to the operator at landfill closure, mostly of sandy to loamy 104

texture, cover thickness varied between 0.1 m and > 2 m. The vegetation that established since 105

the placement of the soil cover consisted of grass, shrubs and small trees. None of the sites had a 106

surface sealing so that landfill gas migrated freely from the waste body through the soil cover. 107

Potential methane oxidation rates as determined in laboratory ranged between 25 and > 2000 g 108

CH4 m-2 d-1 (Gebert et al., 2016). For the purpose of the MiMethox project, all landfills were

(8)

equipped with gas probes, temperature and moisture sensors in 5, 15, 40, and 80 cm depth of the 110

soil cover on three locations per landfill (details in Gebert et al., 2011a). These locations had 111

been chosen based on a preliminary soil and soil gas survey. This study used the same locations. 112

The composition of the methanotrophic community was dominated by Methylocystis, 113

Methylobacter and Methylococcus species among type II, Ia and Ib methanotrophs, respectively,

114

on all of the five investigated landfills (Gebert et al., 2009). Detailed information on the landfills, 115

on the properties of the cover soils, and on the gas fluxes through and from the soil cover is 116

given in Rachor (2012). 117

Further measurements were performed on two methane oxidation cover test cells (cells C and G) 118

on a Dutch landfill (landfill W), constructed with a loamy sand (topsoil) above a loam (subsoil; 119

classification according to FAO, 2006) and a forced gas load to the base of the methane 120

oxidation layer of up to 41 g CH4 m-2 d-1. The test cells had a size of 1.060 m2 each. They served

121

to investigate the combination methane oxidation and measures to limit drainage water 122

infiltration. The latter was realized by a capillary barrier (cell C) and a drainage mat (cell G). 123

Short grass vegetation was maintained on these test cells by regular mowing. More information 124

on the test cell setup is given in (Röwer, 2016a; Geck et al., 2016). 125

2.2 Gas push-pull tests

126

The application of the gas push-pull test to quantify methane oxidation rates in landfill cover 127

soils including a detailed description of the methodology and the nature and interpretation of 128

results on 50 GPPTs is described in depth in Streese-Kleeberg et al. (2011). This section 129

summarizes the principal approach of the GPPT. A gas mixture of 30 l, containing methane, 130

oxygen and argon (10% CH4, 20% Ar in air), were injected at the designated soil depth (usually

(9)

between 40 and 60 cm below surface) at a flow rate of 2.5 l/min using a pump connected to a 132

mass flow meter and a tube with a perforated tip. The depth was chosen so that the methane 133

background concentration would not exceed 5% and the oxygen background concentration 134

indicated oxidative conditions in the soil. Hereby, it was prevented that a high methane 135

background confounded the GPPT results. Soil gas composition was verified beforehand by on-136

site analysis of the soil gas phase using a biogas analyser (BM2K2-E000, Geotechnical 137

Instruments (UK) Ltd.) for measurement of CH4 (infrared detector), CO2 (infrared detector) and

138

O2 (electrochemical detector). The detection limit for all three gases was 0.1 vol.%, Instrument

139

calibration was performed with certified calibration gases in the laboratory and validated with 140

parallel gas-chromatographic analyses. 141

Following injection, the mixture of reactant, tracer and soil gas were extracted from the same 142

tube. The extraction flow rate was 0.3 to 0.5 l/min. During the extraction phase the gas was 143

streamed through 40 ml glass vials with septum caps which were exchanged every 1 to 4 144

minutes. The time interval depended on the rate of decline of the methane concentration. This 145

was continuously measured in the extracted gas stream with an infrared cell. These samples were 146

analysed for methane and argon concentrations. The vials contained a barrier solution (200 g 147

NaCl l-1 and 5 g citric acid l-1 in water) to prevent gaseous diffusion through the vial septa and 148

were stored in an inverted position until analysis. In addition, gas samples intended for carbon 149

stable isotope analyses were collected from the gas stream using a syringe and injected into vials 150

containing a supersaturated salt solution. Supersaturation served to prevent dissolution of gas 151

into the liquid phase. Vials were also stored in an inverted position until stable isotope analyses 152

took place. Three to eleven stable isotope samples were retrieved during each GPPT. The 153

average δ13

C signature of the injected gas was -41.56 ± 0.89‰ (standard deviation). 154

(10)

Data evaluation of the GPPT was performed according to Yang et al. (2007). First, reactant and 155

tracer were corrected for background concentrations for each sampling point. The corrected 156

concentrations were then normalized to the concentration of the injected reactant and tracer. 157

Thereafter, the normalized concentrations were plotted over time elapsed since the 158

commencement of the extraction phase. From the normalized tracer concentrations, the mixing 159

factor was obtained for each sampling point as follows: 160

𝑓𝑓 = (Ctr− Ctrb)/(Ctr0 − Ctrb) Eq. 4

161

Where 𝑓𝑓 = mixing factor 162

Ctr = current tracer concentration

163

Ctrb = background tracer concentration

164

Ctr0 = injected tracer concentration.

165

All concentrations are to be expressed in the same unit. 166

The normalized concentrations decrease at a higher rate than those of the tracer and so-called net 167

mass transfer takes place if the reactant is consumed, i.e. in this case is microbially degraded. Net 168

mass transfer thus represents the change in reactant (methane) concentration that cannot be 169

attributed to dilution and therefore represents methane consumption. The decrease in 170

concentration is due to dispersion in the soil only if the slopes of reactant and tracer are similar. 171

The net mass transfer is calculated as follows: 172

𝑁𝑁𝑁𝑁𝑁𝑁 = 𝐶𝐶𝑟𝑟 – [𝑓𝑓 × 𝐶𝐶𝑟𝑟0+ (1 − 𝑓𝑓) × 𝐶𝐶𝑟𝑟𝑏𝑏] Eq. 5

(11)

Where 𝑁𝑁𝑁𝑁𝑁𝑁 = net mass transfer 174

𝐶𝐶𝑟𝑟 = current reactive gas concentration

175

𝑓𝑓 = mixing factor 176

𝐶𝐶𝑟𝑟0= injected reactive gas concentration

177

𝐶𝐶𝑟𝑟𝑏𝑏 = background reactive gas concentration.

178

All concentrations are to be expressed in the same unit, the net mass transfer then assumes this 179

unit. 180

Prerequisite for the validity of the approach is that reactant and tracer gases exhibit similar 181

transport characteristics in the soil gas phase. In the case of methane, argon meets this condition 182

(Gómez et al., 2009). The methane oxidation rate is derived as a zero order kinetic constant from 183

the slope of the net mass transfer over time: 184

𝑘𝑘0 = 𝑁𝑁𝑁𝑁𝑁𝑁/𝑡𝑡 Eq. 6

185

Where 𝑘𝑘0 = zero order rate constant and 𝑡𝑡 = time. 186

The zero order rate constant yields the methane oxidation rate in mass unit CH4 per volume unit

187

soil air per unit time. It has to be considered that this rate is to be interpreted as a potential rate, 188

as both oxygen and methane are supplied at optimum concentration levels during the test. 189

2.3 Analysis of stable isotopes

190

The δ13C ratio was analysed in triplicate using a GC-IRMS (gas chromatography isotope ratio 191

mass spectrometry, Delta Plus, ThermoScientific, Dreieich, Germany) equipped with a 25 m 192

(12)

capillary column (Poraplot, 0.32 mm). The reference standard NGS3 8561 (δ13CVPDB-NGS3 193

8561 = -73.27‰; NIST, Gaithersburg, MD, USA) was used to express the methane-derived δ13C 194

ratio relative to the VPDB (Vienna Pee Dee Belemnite) standard. Analytical replicate precision 195

was < 0.2‰. 196

2.4 Calculation of fractionation factor αox 197

Values for αox were obtained using GPPT and methane-derived δ13C-data based on the approach

198

derived by Coleman et al. (1981) from Rayleigh (1896): 199 𝛿𝛿13𝐶𝐶 𝑡𝑡≅ 1000 ∗ �1𝑎𝑎𝑎𝑎− 1� × ln �𝑀𝑀𝑀𝑀0� + 𝛿𝛿13𝐶𝐶𝑡𝑡=0 Eq. 7 200 Where 201 δ13

Ct = δ13C value of CH4 remaining at time t at the particular soil depth during the GPPT

202

M/M0 = fraction of CH4 (corrected for dilution) remaining at time t at the particular soil

203

depth related to the initial methane concentration M0 (also corrected for dilution)

204

δ13

Ct=0 = δ13C value of CH4 at the initial time, i.e. of the mixture injected in to the soil.

205

The slope of the line δ13

Ct - δ13Ct=0 versus ln(M/M0) (see also Figure 2) is equal to 1000*(1/αox

-206

αtrans), so that αox was calculated as follows:

207 1 𝛼𝛼𝑎𝑎𝑎𝑎 = 1 + �𝛿𝛿13𝐶𝐶𝑡𝑡−𝛿𝛿13𝐶𝐶𝑡𝑡=0 ln�𝑀𝑀0𝑀𝑀�×1000 Eq. 8 208

Pushing and pulling the gas into and out of the soil results in advective gas transport. Methane 209

concentrations can be assumed to be spatially uniform within the injected and extracted volume a 210

(13)

given point in time. Hence, gas transport driven by a concentration gradient was assumed to be 211

insignificant within the time frame of the test. Therefore, fractionation due to diffusive gas 212

transport (Barker and Fritz, 1981) was neglected and αtrans was assumed to be equal to 1. Values

213

for αox were only accepted when the correlation coefficient of the slope was significant on a 99%

214

confidence level. 215

4 RESULTS AND DISCUSSION

216

Table 1 summarizes the methane oxidation rates derived from the GPPTs, the details of the 217

respective environmental conditions, the soil properties as well as the calculated values αox.

218

Some of the information can also be found in Streese-Kleeberg et al. (2011) who introduce the 219

GPPT method adapted for landfill cover soils and present data on more than 50 GPPTs. The 220

following sections discuss the data on environmental conditions, soil properties, and the 221

variability of αox in relation to the oxidation rate.

222

Table 1: Soil properties, environmental conditions, fractionation factors and oxidation rates for each site. n.d. 223

= not determined. Temp. = temperature, moist. = moisture. Site ID notation: H, K, D, L, A = different MSW 224

landfills in north western Germany; WC, WG = test cell C and test cell G on landfill W in The Netherlands. 225

P1-5 = individual sampling points. 226

4.1 Environmental conditions, soil properties and methane oxidation potential

227

Both the environmental conditions and soil properties such as porosity varied between the 228

individual points of investigation but also for the same point over time. Over all landfill sites and 229

points of investigation, soil temperature spanned between 1.0 and 20 °C, and soil moisture 230

ranged between 3.5 and 26.1 vol%. Air-filled porosity, predominantly determining the soil’s gas 231

transport properties, varied between 10.2 and 32.9 vol%. 232

(14)

The seasonal variation of soil temperature and moisture influences the microbial process of 233

methane oxidation (e.g. Gebert et al., 2003; Spokas and Bogner, 2011) and hence also emission 234

(e.g. Tecle et al., 2009; Rachor et al., 2013; Geck et al., 2016). The variability of texture, 235

compaction and moisture strongly impacts soil gas diffusivity and gas conductivity (Møldrup et 236

al., 2000; Gebhardt et al., 2009; Gebert et al., 2011b; Röwer et al., 2016b) and thereby modulates 237

the methane oxidation rate in landfill soil covers by influencing the ingress of atmospheric air. 238

Further variability of methane oxidation rates is introduced by the varying level of exposure of 239

the methanotrophic community to methane (Röwer et al., 2011; Schroth et al., 2012), leading to 240

methane oxidation rates spanning several orders of magnitude on one or between different 241

landfill sites (Spokas and Bogner, 2011; Gebert et al., 2016). 242

Given the above, significant spatial and temporal variability of the methane oxidation rate in 243

landfill cover soils can be expected a priori and was also observed during the gas push-pull tests 244

conducted in this study (Table 1). For the six investigated landfill cover soils the range of 245

oxidation rates found with the GPPTs ranged between 2.3 and 144 g m3soil air-1 h-1, corresponding

246

to 0.32 and 18 g m3soil-1 h-1 when values for total porosity and soil moisture at the time of

247

measurement were considered. These oxidation rates relate to efficiencies of 3% and 100% at the 248

end of the GPPT. An extraordinarily high rate of 1110 g m3soil air-1 h-1 or 354 g m3soil-1 h-1 was

249

found for location WC1. Sites WC1 and WC4 are located within 10 m of each other on a test cell 250

investigating the suitability of biocovers for the mitigation of landfill methane. As indicated by 251

gas profile measurements at both sites (data not shown), WC1 receives high concentrations of 252

methane whereas due to spatial heterogeneity of gas fluxes the exposure to landfill gas at WC4 is 253

low. As a result, WC1 developed a very high potential activity of 1110 g CH4 m-3soil air h-1,

254

compared to only 50 g CH4 m-3soil air h-1 at WC4. Evolution of methanotrophic potential strongly

(15)

depends on the supply of the substrate methane (Röwer et al., 2011; Spokas and Bogner, 2011; 256

Schroth et al., 2012) and can hence span orders of magnitude in relation to the spatial 257

heterogeneity of fluxes. In line with this, GPPT-based oxidation potentials for the cover soil on a 258

Swiss municipal solid waste landfill were found to be as high as 210 g m3soil-1 h-1 for sites where

259

high methane concentrations were measured in the soil gas phase. The oxidation potential was 260

below the GPPT’s detection limit where soil methane concentrations were very low (Schroth et 261

al., 2012). 262

4.2 Shift of the methane δ13C signature with oxidation efficiency

263

During the time course of each GPPT, the δ13C signature in the sampled soil gas phase shifted 264

towards less negative values, i.e. the heavier isotopologue 13CH4 became enriched as the lighter

265

isotopologue (12CH4) was preferentially consumed by the methanotrophic bacteria. In Figure 1,

266

the change in isotopic signature is plotted versus the oxidation efficiency (fox) determined for

267

each individual sampling point during each of the 22 GPPTs. For the entire data set, the 268

efficiency ranged between near 0% and near 100%. The maximum observed enrichment in δ13C 269

was 62‰ (from -41 to +21‰ for GPPT no. 26 at site HP5, compare also Table 1), with large 270

shifts suggesting vigorous methane oxidation activity. As expected, the degree of enrichment 271

generally increased with increasing oxidation efficiency. However, some data points showed 272

little enrichment in relation to the oxidation efficiency while for others, the extent of enrichment 273

was substantially higher than would be expected from the relationship between fox and δ13C for

274

the bulk data, indicating different extents of isotopic fractionation during the oxidation process. 275

Figure 1: Relationship between methane oxidation efficiency fox and respective shift in carbon isotopic 276

signature of methane relative to the injected gas. Data from 22 individual GPPTs with three to six individual 277

points per GPPT (compare also Table 1). 278

(16)

Figure 2 shows how the change in isotopic signature over time (δ13Ct - δ13Ct=0) relates to the

279

natural logarithm of the remaining methane (ln(M/M0)), using an example of a GPPT yielding a

280

low oxidation rate (#51, site DP1) and one yielding a high oxidation rate (#59, also site DP1). 281

According to Eq. 8, values for αox were calculated from the slope of this line, if significant on a

282

99% level. 283

Figure 2: Relationship between the fraction of methane remaining during the course of a GPPT, expressed as 284

ln(M/M0), and the change in isotopic signature, expressed as δ13Ct - δ13Ct=0, for GPPTs 51 (left panel, point 285

DP1) and 59 (right panel, also point DP1). Line = linear fit. 286

4.3 Calculated values for αox and their relationship to the methane oxidation rate 287

The literature reports values for αox between 1.018 and 1.049 (Chanton et al., 1999; Chanton and

288

Liptay, 2000; Börjesson et al., 2001, 2007; De Visscher et al., 2004; Capanema and Cabral, 289

2012), a range also represented by 64% of the measurements in this study (Table 1, Figure 3). 290

Two GPPTs showed lower and six of the 22 GPPTs higher isotopic fractionation. The 291

relationship between the in-situ methane oxidation rate derived from the GPPTs and the values 292

for αox is shown in Figure 3. Values for αox were high at low oxidation rates and decreased

293

nonlinearly with increasing methane oxidation rate, approximating the value of αox = 1 (i.e. no

294

fractionation) at very high rates (sampling point WC1). Hardly any discrimination of the heavier 295

isotopologue was observed under these conditions of high methanotrophic activity (GPPTs #53 296

at site KP2, #103 at site WC3). Vice versa, a low activity allows for a high extent of 297

discrimination (e.g. GPPTs #41 at site DP1, #44 at site LP1). The observed fractionation factor 298

not only varied between the different sites but also for different GPPTs conducted at the same 299

landfill and the same sampling point, presumably reflecting the temporal variability of process 300

controls (e.g. temperature, moisture and hence oxygen supply) over the seasons. The maximum 301

(17)

span of values for αox observed for one individual point of investigation was 1.035 to 1.100

302

(point DP1, Table 1, Figure 3). 303

Figure 3: Relationship between methane oxidation rate and fractionation factor αox. n = 22. For site IDs see

304

heading of Table 1. 305

The general relationship between environmental conditions, the methane oxidation rate and the 306

fractionation factor αox can be deduced from the largest single data set available from site DP1 (n

307

= 7), shown in Figure 4. It is seen that cooler conditions in the winter coincide with increased 308

soil moisture, which is typical for the temperate European climate with the lower methane 309

oxidation potentials and wit higher values for αox. Vice versa, the highest oxidation potentials

310

and the lowest extent of fractionation are found when the soil is warmer and dryer. The same 311

seasonal pattern is apparent for sited HP5 and KP2. All in all, the individual data sets are too 312

small for a multivariate statistical analysis of the factors of influence, which, however, have been 313

previously described (e.g. Scheutz et al., 2009). 314

Figure 4: Relationship between environmental parameters, methane oxidation rate and fractionation factor 315

αox as found at site DP1 (data from Table 1).

316

The data from sampling point DP1 were also used to assess the effect of the variability of αox on

317

calculated values for the oxidation efficiency fox.by example of GPPT #28. To this end, fox was

318

calculated for the change in isotopic signature observed for this particular test (40.36‰ to -319

16.54‰) employing the seven individual values for αox determined during the seven GPPTs

320

conducted at site DP1 (1.036 to 1.100, see Table 1). The resulting oxidation efficiencies varied 321

between -2.5% and 66%, indicating that it is not possible to assume αox to be constant even in the

322

very same sampling location (in this case, DP1). 323

In another sensitivity analysis, Capanema and Cabral (2012) varied the value of αox by ±0.5% of

324

the value determined in the laboratory on their own samples. The resulting methane oxidation 325

(18)

efficiencies (fox) varied between 53 and 88% for the same data set. Increasing αox by 0.5% led to

326

implausible oxidation efficiencies greater than 100%. All examples show that the sensitivity of 327

the method to the correct value of αox is hence very high and that hence the exact value of αox

328

needs to be determined for each individual combination of conditions under which the methane 329

oxidation efficiency of a landfill cover is assessed. 330

The observed variation of isotopic fractionation with oxidation rate corroborates findings from 331

earlier studies. Templeton et al. (2006) showed that most important factor determining 332

fractionation of methane isotopologues was the fraction of total methane oxidized per time unit, 333

with values for αox decreasing with increasing turnover rates and increasing cell densities.

334

Chanton et al. (2008) found αox to decline nonlinearly with increasing maximum methane

335

oxidation rates (vmax), indicating less discrimination at higher turnover rates when the

336

temperature is constant. 337

In summary, it was shown that values for αox varied at the same sampling point, between

338

different sampling points on the same landfill, and between different landfills. Further, the 339

oxidation rate strongly affected the extent of discrimination of the heavy isotopologue and hence 340

the magnitude of αox. Low rates caused a high extent of fractionation whereas at very high rates

341

αox approximated the value of one, i.e. no preferential oxidation of the lighter isotopologue was

342

observed. In conclusion, the exact value for αox and its variability in time and space should be

343

known to reduce uncertainty in the determination of the methane oxidation efficiency. Since soil 344

temperature and moisture as well as the spatial and temporal heterogeneity of gas fluxes through 345

landfill covers result in continuously changing environmental conditions, the extent of both 346

transport and biological fractionation will vary correspondingly. Adopting αox values from the

347

literature or assigning the same α value to the same measurement point over time or assigning 348

(19)

the same value to different points on one or several landfills can lead to misjudgment of methane 349

oxidation efficiencies. In order to avoid this, αox would have to be determined for each individual

350

campaign using undisturbed soils samples at the prevailing moisture content. Depending on the 351

desired level of accuracy, these limitations affect the applicability of the stable isotope approach 352

as a robust, practical and economic method for the exact quantification of methane oxidation 353

efficiencies. 354

FUNDING SOURCES 355

The research was funded by the German Federal Ministry for Education and Research (BMBF) 356

within the framework of the project MiMethox (Microbial Methane Oxidation in Landfill 357

Covers). 358

ACKNOWLEDGMENT 359

The authors would like to thank Dr. Ingke Rachor, Dr. Kim-Karen Kleeberg, Volker 360

Kleinschmidt, Cindy Streblow, Simon Klauke and Dr. Christian Knoblauch for their assistance 361

in field and laboratory measurements. 362

(20)

REFERENCES 363

Abichou, T., Mahieu, K., Chanton, J., Romdhane, M., Mansouri, I. (2011): Scaling methane 364

oxidation: From laboratory incubation experiments to landfill cover field conditions. Waste 365

Management 31, 978-986. 366

Barker, J.F., Fritz, P. (1981): Carbon isotope fractionation during microbial methane oxidation. 367

Nature 293, 289-291. 368

Blair, N., Leu, A., Munoz, E., Olson, J., Kwong, E., Des Marais, D. (1985): Carbon isotopic 369

fractionation in heterotrophic microbial metabolism. Applied and Environmental Microbiology 370

50, 996-1001. 371

Bogner J., Abdelrafie Ahmed, M., Diaz, C., Faaij, A., Gao, Q., Hashimoto, S., Mareckova, K., 372

Pipatti, R., Zhang, T. (2007): Waste Management. In: Climate Change 2007: Mitigation. 373

Contribution of Working Group III to the Fourth Assessment Report of the Intergovernmental 374

Panel on Climate Change. Metz, B., Davidson, O.R., Bosch, P.R., Dave, R., Meyer, L.A. (eds). 375

Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA. 376

Börjesson, G., Chanton, J., Svensson, B. (2001): Methane oxidation in two Swedish landfill 377

covers measured with carbon-13 to carbon-12 isotope ratios. Journal of Environmental Quality 378

30, 369-376. 379

Börjesson, G., Samuelsson, J., Chanton, J. (2007): Methane oxidation in Swedish landfills 380

quantified with the stable carbon isotope technique in combination with an optical method for 381

emitted methane. Environmental Science & Technology 41, 6684-6690. 382

Cabral, A.R., Capanema, M.A., Gebert, J., Moreira, J.F., Jugnia, L.B. (2009): Quantifying 383

microbial methane oxidation efficiencies in two experimental landfill biocovers using stable 384

(21)

isotopes. Water, Air, and Soil Pollution 209, 157-172. 385

Capanema, M.A., Cabral, A.R. (2012): Evaluating methane oxidation efficiencies in 386

experimental landfill biocovers by mass balance and carbon stable isotopes. Water, Air, and Soil 387

Pollution 223, 5623-5635. 388

Chanton, J.P., Liptay, K. (2000): Seasonal variation in methane oxidation in a landfill cover soil 389

as determined by an in situ stable isotope technique. Global Biogeochemical Cycles 14, 51-60. 390

Chanton, J.P., Powelson, D.K., Abichou, T., Fields, D., Green, R. (2008): Effect of temperature 391

and oxidation rate on carbon-isotope fractionation during methane oxidation by landfill cover 392

materials. Environmental Science & Technology 42, 7818-7823. 393

Chanton, J.P., Rutkowski, C.M., Mosher, B. (1999): Quantifying methane oxidation from 394

landfills using stable isotope analysis of downwind plumes. Environmental Science & 395

Technology 33, 3755-3760. 396

Chanton, J., Abichou, T., Langford, K., Spokas, K., Hater, G., Green, R., Goldsmith, D., Barlaz, 397

M.A. (2011): Observations on the methane oxidation capacity of landfill soils. Waste 398

Management 31, 914-925. 399

Coleman, D.D., Risatti, J.B. (1981): Fractionation of carbon and hydrogen isotopes by methane-400

oxidizing bacteria. Geochimica et Cosmochimica Acta 45, 1033-1037. 401

De Visscher, A., De Pourq, I., Chanton, J. (2004): Isotope fractionation effects by diffusion and 402

methane oxidation in landfill cover soils. Journal of Geophysical Research 109, (D18111). 403

European Environment Agency (EEA). Annual European Community Greenhouse Gas 404

Inventory 1990-2007 and Inventory Report 2007. Submission to the UNFCCC Secretariat. EEA 405

(22)

Technical report, No. 4. 2009; Annex Key Category Analysis. 406

http://reports.eea.europa.eu/technical_report_2006_6/en. 407

FAO - ISRIC & ISSS (eds.) (2006): World reference base for soil resources 2006: a framework 408

for international classification, correlation and communication. World soil resources reports, 409

103. Rome. 410

Gebert, J., Rachor, I., Gröngröft, A., Pfeiffer, E.-M. (2011a): Temporal variability of soil gas 411

composition in landfill covers. Waste Management 31, 935-945. 412

Gebert, J., Gröngröft, A., Pfeiffer, E.-M. (2011b): Relevance of soil physical properties for the 413

microbial oxidation of methane in landfill covers. Soil Biology and Biochemistry 43, 1759-1767. 414

Gebert, J., Singh, B.K., Pan, Y., Bodrossy, L. (2009): Activity and structure of methanotrophic 415

communities in landfill cover soils. Environmental Microbiology Reports 1, 414-423. 416

Gebert, J, Rachor, I.M., Streese-Kleeberg, J., Pfeiffer, E.-M. (2016): Methane oxidation in a 417

landfill cover soil under conditions of diffusive and advective flux, assessed by in-situ and ex-418

situ methods. Current Environmental Engineering, 3, 144-160. 419

Gebhardt, S., Fleige, H., Horn, R. (2009): Effect of compaction on pore functions of soils in a 420

Saalean moraine landscape in North Germany. Journal of Plant Nutrition and Soil Science 172, 421

688-695. 422

Geck, C., Scharff, H., Pfeiffer, E.-M., Gebert, J. (2016): Validation of a simple model to predict 423

the performance of methane oxidation systems, using field data from a large scale biocover test 424

field. Waste Management 56, 280-289. 425

(23)

Gómez, K.E., Gonzalez-Gil, G., Lazzaro, A., Schroth, M.H. (2009): Quantifying methane 426

oxidation in a landfill-cover soil by gas push-pull tests. Waste Management 29, 2518-2526. 427

Huber-Humer, M., Gebert, J., Hilger, H. (2008): Biotic systems to mitigate landfill methane 428

emissions. Waste Management & Research 26, 33-46. 429

Jahnke, L.L., Summons, R.E., Hope, J.M., Des Marais, D.J. (1999): Carbon isotopic 430

fractionation in lipids from methanotrophic bacteria II: The effects of physiology and 431

environmental parameters on the biosynthesis and isotopic signatures of biomarkers. Geochimica 432

et Cosmochimica Acta 63, 79-93. 433

Møldrup, P., Olesen, T., Schjønning, P., Yamaguchi, T., Rolston, D.E. (2000): Predicting the gas 434

diffusion coefficient in undisturbed soil from soil water characteristics. Soil Science Society of 435

America Journal 64, 1588-1594. 436

Rachor, I.M. (2012): Spatial and temporal patterns of methane fluxes on old landfills: processes 437

and emission reduction potential. Dissertation. Hamburger Bodenkundliche Arbeiten 67. ISSN 438

0724-6382. 439

Rachor, I.M., Streese-Kleeberg, J., Gebert, J. (2009): Spatial and temporal variability of gas 440

emissions from old landfills. Proceedings Sardinia, Twelfth International Waste Management 441

and Landfill Symposium S. Margherita di Pula, Cagliari, Italy; 5 - 9 October 2009. 2009 by 442

CISA, Environmental Sanitary Engineering Centre, Italy. 443

Rachor, I.M., Gebert, J., Gröngröft, A., Pfeiffer, E.-M. (2013): Variability of methane emissions 444

from an old landfill over different time-scales. European Journal of Soil Science 64, 16-26. 445

Rayleigh, J.W.S. (1896): Theoretical considerations respecting the separation of gases by 446

diffusion and similar processes. Philosophical Magazine 42, 493-498. 447

(24)

Reeburgh, W.S., Hirsch, A.I., Sansone, F.J., Popp, B.N., Rust, T.M. (1997): Carbon kinetic 448

isotope effect accompanying microbial oxidation of methane in boreal forest soils. Geochimica 449

et Cosmochimica Acta 61, 4761-4767. 450

Röwer, I.U., Geck, C., Gebert, J., Pfeiffer, E.-M. (2011): Spatial variability of soil gas 451

concentrations and methane oxidation in landfill cover soils. Waste Management 31, 926-934. 452

Röwer, I.U., Scharff, H., Pfeiffer, E.-M., Gebert, J. (2016a): Optimized landfill biocover for CH4

453

oxidation I: Experimental design and oxidation performance. Current Environmental 454

Engineering 3, 80-93. 455

Röwer, I.U., Streese-Kleeberg, J., Scharff, H., Pfeiffer, E.-M., Gebert, J. (2016b): Optimized 456

landfill biocover for CH4 oxidation II: Implications of spatially heterogeneous fluxes for

457

monitoring and emission prediction. Current Environmental Engineering 3, 94-106. 458

Scheutz, C., Kjeldsen, P., Bogner, J.E., De Visscher, A., Gebert, J., Hilger, H. A., Huber-Humer, 459

M., Spokas, K. (2009): Microbial methane oxidation processes and technologies for mitigation of 460

landfill gas emissions. Waste Management & Research 27, 409-455. 461

Schroth, M.H., Eugster, W., Gómez, K.E., Gonzalez-Gil, G., Niklaus, O.A., Oester, P. (2012): 462

Above- and below-ground methane fluxes and methanotrophic activity in a landfill-cover soil. 463

Waste Management 32, 879-889. 464

Spokas, K.A., Bogner, J.E. (2011): Limits and dynamics of methane oxidation in landfill cover 465

soils. Waste Management, 31, 823-832. 466

Streese-Kleeberg, J., Rachor, I., Gebert, J., Stegmann, R. (2011): Use of gas push-pull tests for 467

the measurement of methane oxidation in different landfill cover soils. Waste Management 31, 468

995-1001. 469

(25)

Tecle, D., Lee, J., Hasan, S. (2009): Quantitative analysis of physical and geotechnical factors 470

affecting methane emission in municipal solid waste landfill. Environmental Geology 56, 1135-471

1143. 472

Templeton, A.S., Chu, K.-H., Alvarez-Cohen, L., Conrad, M.E. (2006): Variable carbon isotope 473

fractionation expressed by aerobic CH4-oxidizing bacteria. Geochimica et Cosmochimica Acta

474

70, 1739-1752. 475

Tyler, S.C., Crill P.M., Brailsford, G.W. (1994): 13C/12C Fractionation of methane during 476

oxidation in a temperate forested soil. Geochimica et Cosmochimica Acta 58, 1625-1633. 477

Urmann, K., Gonzalez-Gil, G., Schroth, M.H., Hofer, M., Zeyer, J. (2005): New field method: 478

gas push-pull test for in situ quantification of microbial activities in the vadose zone. 479

Environmental Science & Technology 39, 304-310. 480

United States Environmental Protection Agency (USEPA), 2016: Inventory of U.S. Greenhouse 481

Gas Emissions and Sinks: 1990-2014. EPA 430-R-16-002. 482

Widory, D., Proust, E., Bellenfant, G., Bour, O. (2012): Assessing methane oxidation under 483

landfill covers and its contribution to the above atmospheric CO2 levels: the added value of the

484

isotope (δ13C and δ18

O CO2; δ13C and δD CH4) approach. Waste Management 32, 1685-1692.

485

Yang, C.B., Park, M., Zhu, C. (2007): A method for estimating in situ reaction rates from push-486

pull experiments for arbitrary solute background concentrations. Environmental & Engineering 487

Geoscience 13, 345-354. 488

(26)

List of figure captions

489

Figure 1: Relationship between methane oxidation efficiency fox and respective shift in carbon

490

isotopic signature of methane relative to the injected gas. Data from 22 individual GPPTs with 491

three to six individual points per GPPT (compare also Table 1). 492

Figure 2: Relationship between the fraction of methane remaining during the course of a GPPT, 493

expressed as ln(M/M0), and the change in isotopic signature, expressed as δ13

Ct - δ13Ct=0, for 494

GPPTs 51 (left panel, point DP1) and 59 (right panel, also point DP1). Line = linear fit. 495

Figure 3: Relationship between methane oxidation rate and fractionation factor αox. n = 22. For

496

site IDs see heading of Table 1. 497

Figure 4: Relationship between environmental parameters, methane oxidation rate and 498

fractionation factor αox as found at site DP1 (data from Table 1).

499

List of table captions

500

Table 1: Soil properties, environmental conditions, fractionation factors and oxidation rates for 501

each site. n.d. = not determined. Temp. = temperature, moist. = moisture. Site ID notation: H, K, 502

D, L, A = different MSW landfills in north western Germany; WC, WG = test cell C and test cell 503

G on landfill W in The Netherlands. P1-5 = individual sampling points. 504

(27)

Figures

505

506

Figure 1: Relationship between methane oxidation efficiency fox and respective shift in carbon isotopic

507

signature of methane relative to the injected gas. Data from 22 individual GPPTs with three to six individual 508

points per GPPT (compare also Table 1). 509 0 10 20 30 40 50 60 70 80 90 100 0 5 10 15 20 25 30 35 40 45 50 55 60 65

S

h

ift i

n

δ

13

C

s

ignat

ur

e [

]

Oxidation efficiency f

ox

[%]

(28)

510

Figure 2: Relationship between the fraction of methane remaining during the course of a GPPT, expressed as 511

ln(M/M0), and the change in isotopic signature, expressed as δ13Ct - δ13Ct=0, for GPPTs 51 (left panel, point 512

DP1) and 59 (right panel, also point DP1). Line = linear fit. 513 -0.14 -0.12 -0.10 -0.08 -0.06 -0.04 -0.02 0.00 0 1 2 3 4 5 6 7 8 9 δt -δt= 0 [‰ ] ln(M/M0) [v/v] GPPT 51 slope = -46.70 r = -0.995 Oxidation rate 3.6 g m-3 soil air h -1 -0.5 -0.4 -0.3 -0.2 -0.1 0.0 0 5 10 15 20 25 30 Oxidation rate 46.2 g m-3soil air h-1 δt -δt= 0 [‰ ] ln(M/M0) [v/v] GPPT 59 slope = -56.87 r = -0.997

(29)

0 50 100 150 200 250 300 350 1000 1200 1,00 1,02 1,04 1,06 1,08 1,10 1,12 1,14 1,16 Site ID DP1 AP1 HP5 KP2 LP1 WC1 WG2 WC4 KP3

Frac

ti

onat

ion f

ac

tor

a

ox

[-]

CH

4

oxidation rate [g CH

4

m

-3soil air

h

-1

]

514

Figure 3: Relationship between methane oxidation rate and fractionation factor αox. n = 22. For site IDs see

515

heading of Table 1. 516

(30)

517

Figure 4: Relationship between environmental parameters, methane oxidation rate and fractionation factor 518

αox as found at site DP1 (data from Table 1).

519 0 5 10 15 20 So il t em per atu re (° C) 0 5 10 15 20 Soi l m oi st ur e (v o. %) 0 10 20 30 40 50 CH4 ox id at io n p ot en tia l (g m -3 so il a ir h -1) 1.00 1.02 1.04 1.06 1.08 1.10 1.12 01. 10. 2008 31. 10. 2008 01. 12. 2008 31. 12. 2008 31. 01. 2009 02. 03. 2009 02. 04. 2009 02. 05. 2009 02. 06. 2009 Fr act io na tio n f act or a lpha _o x

(31)

Tables

520

Table 1: Soil properties, environmental conditions, fractionation factors and oxidation rates 521

(potentials) for each site. n.d. = not determined. Temp. = temperature, moist. = moisture. Site ID 522

notation: H, K, D, L, A = different MSW landfills in north western Germany; WC, WG = test 523

cell C and test cell G on landfill W in The Netherlands. P1-5 = individual sampling points. 524 GPPT No. Site ID Date Soil temp. Total porosity Soil moist. Air-filled porosity Total shift in δ13 C Background CH4 Background O2 αox CH4 oxidation rate (°C) (vol.%) (vol.%) (vol.%) (‰) (vol.%) (vol.%) (-) (g m-3soil air h-1)

26 HP5 13.10.08 14.9 51.0 24.4 26.6 62.04 0.0 15.0 1.020 91.3 38 HP5 02.02.09 1.0 51.0 23.4 27.6 34.00 0.22 3.1 1.019 50.6 46 HP5 17.03.09 8.1 42.8 24.4 18.4 15.62 0.0 6.6 1.096 6.50 62 HP5 30.06.09 18.0 51.0 24.5 26.5 43.18 0.0 19.1 1.026 104 65 HP5 03.08.09 20.0 51.0 24.9 26.1 23.78 0.0 19.3 1.033 90.9 27 KP2 20.10.08 12.6 34.5 24.3 10.2 52.68 0.0 5.4 1.018 66.5 29 KP2 03.11.08 10.2 34.5 23.3 11.2 36.63 0.0 8.9 1.023 41.9 53 KP2 27.04.09 12.8 34.5 22.0 12.5 17.44 0.24 16.1 1.005 143.5 28 DP1 29.10.08 15.4 42.8 10.3 32.5 23.81 0.008 11.0 1.036 34.2 31 DP1 24.11.08 10.2 42.8 10.0 32.8 24.51 0.0 7.5 1.035 5.50 34 DP1 16.12.08 8.7 42.8 9.9 32.9 20.46 0.0 16.0 1.064 10.0 41 DP1 17.02.09 4.9 42.8 18.2 24.6 11.86 0.0 4.8 1.100 4.30 51 DP1 20.04.09 12.6 42.8 14.9 27.9 8.13 0.0 10.5 1.049 3.56 55 DP1 26.05.09 16.5 42.8 12.4 30.4 25.95 0.0 12.2 1.044 16.4 59 DP1 23.06.09 17.3 42.8 11.5 31.3 26.25 0.0 9.8 1.060 46.2 32 AP1 25.11.08 4.5 n.d. 21.1 n.d. 13.21 0.47 0.6 1.042 3.20 60 AP1 24.06.09 16.4 n.d. 13.9 n.d. 23.57 0.0 13.1 1.042 34.7 44 LP1 09.03.09 4.3 39.8 26.1 13.7 4.16 0.0 17.2 1.151 2.30 103 WC1 13.07.11 17.3 49.2 17.3 31.9 34.19 0.74 4.4 1.006 1110 104 WC4 13.07.11 17.0 49.2 17.1 32.1 32.99 0.0 17.0 1.033 49.5 105 WG2 13.07.11 20.0 43.2 17.1 26.1 8.47 0.0 17.7 1.030 17.3 119 KP3 17.09.12 16.2 32.6 3.5 29.1 24.56 0.0 15.0 1.034 69.1 525

Cytaty

Powiązane dokumenty

Życie konsumpcyjne to niekończące się inicjowanie nowości, których potrzebę zapewnia także dwuznaczność komunikatów, poddawanie się konwencji postrzeganej jako

Zaw arte w opasłym (i nadzwyczaj starannie wydanym, podziwu godne są zwłaszcza znakom ite fotografie, co zresztą jest tradycją i poniekąd specjalnością tego wydawnictwa) tom

The primary concern of the study was to compare pro-health behaviors in participants with stable coronary artery disease (SCAD), also known as stable ischemic heart disease (SIHD),

Przy pomocy spektroskopii MIR oraz Ramana określono wpływ modyfi katorów w postaci lantanu i lutetu na strukturę szkieł tellurowych.. Na podstawie badań optycznych transmisji i

Here we study the Brill–Noether theory of “extremal” Cornalba’s theta-characteristics on stable curves C of genus g, where “extremal” means that they are line bundles on

A new model was proposed which is able to produce synthetic marine projects with linked Markov chains based on statistics of the metocean data near the project loca- tion. In this

onanizm u,

Uznaje się powszechnie, że jednym z najbardziej istotnych elem entów każdej, a przynajmniej bardziej zaawansowanej teorii współczesnego przyrodoznawstwa jest