Sunteți pe pagina 1din 9

Water Air Soil Pollut (2014) 225:2216

DOI 10.1007/s11270-014-2216-2

Mapping Methane and Carbon Dioxide Concentrations


and 13C Values in the Atmosphere of Two Australian Coal
Seam Gas Fields
Damien T. Maher & Isaac R. Santos & Douglas R. Tait

Received: 30 June 2014 / Accepted: 5 November 2014


# Springer International Publishing Switzerland 2014

Abstract Fugitive greenhouse gas emissions from unconventional gas extraction processes (e.g. shale gas,
tight gas and coal bed methane/coal seam gas) are
poorly understood due in part to the extensive area over
which these emissions may occur. We apply a rapid
qualitative approach for source assessment at the scale
of a large gas field. A mobile cavity ring down spectrometer (Picarro G2201-i) was used to provide realtime, high-precision methane and carbon dioxide concentration and carbon isotope ratios (13C), allowing for
on the fly decision making and therefore an efficient
and dynamic surveying approach. The system was used
to map the atmosphere of a production coal seam gas
(CSG) field (Tara region, Australia), an area containing
pre-production exploration CSG wells (Casino,
Australia), and various other potential CO2 and CH4
sources (i.e. wetlands, sewage treatment plants, landfills, urban areas and bushfires). Results showed a widespread enrichment of both CH4 (up to 6.89 ppm) and
CO2 (up to 541 ppm) within the production gas field,
compared to outside. The CH4 and CO2 13C source
values showed distinct differences within and outside
the production field, indicating a CH4 source within the
production field that has a 13C signature comparable to
the regional CSG. While this study demonstrates how
the method can be used to qualitatively assess the location and source of emissions, integration with
D. T. Maher (*) : I. R. Santos : D. R. Tait
Centre for Coastal Biogeochemistry Research, School of
Environment, Science and Engineering, Southern Cross
University, Lismore, NSW 2480, Australia
e-mail: damien.maher@scu.edu.au

atmospheric models may allow for quantitative assessment of emissions. The distinct patterns observed within
the CSG field demonstrates the need to fully quantify
the atmospheric flux of natural and anthropogenic, point
and diffuse sources of greenhouse gases from individual
Australian gas fields before and after production
commences.
Keywords Methane . Carbon dioxide . Cavity ringdown
spectroscopy . Natural gas . Coal seam gas . Greenhouse
gas . Fugitive emissions

1 Introduction
The extraction of methane via unconventional coal
seam gas (CSG) and shale gas fields is rapidly
expanding worldwide. The CSG expansion is due to
technology advances (e.g. horizontal drilling and improved hydraulic fracturing techniques) and an associated decrease in the cost of extraction. Unconventional
gas has been proclaimed to be a suitable bridging fuel
in part due to the lower greenhouse gas emissions
thought to be associated with energy production through
gas rather than coal-fired power stations (Hayhoe et al.
2002; Jaramillo et al. 2007). However, the greenhouse
gas (GHG) footprint of energy production through fossil
fuel combustion is larger than simple end-use combustion (Moore et al. 2014). Development of gas fields can
lead to fugitive emissions via various pathways including well construction, production, venting/flaring
and transportation (IPCC 2006). Measurement of

2216, Page 2 of 9

fugitive emissions from unconventional gas fields is


complicated due to the decentralised infrastructure and
large number of well heads. In addition, when techniques such as directional drilling and hydraulic fracturing are used, methane may travel into overlying sediments and groundwater aquifers through existing or
artificial conduits such as faulty well casings (Osborn
et al. 2011). A recent study showing strong correlations
between atmospheric radon and CSG development in an
Australian CSG field suggests that diffuse as well as
point source emissions may occur (Tait et al. 2013). The
magnitude of the atmospheric flux associated with this
diffuse source is currently unknown and difficult to
estimate.
Recent evidence suggests that emissions from unconventional gas fields may be underestimated and as much as
12 % of gas production may be emitted to the atmosphere
through fugitive emissions (Ptron et al. 2012; Howarth
et al. 2011; Karion et al. 2013), although these values have
been disputed (Cathles et al. 2012) and are likely to be site
specific. Current approaches rely mostly on bottom-up
approaches, i.e. upscaling average emissions from individual components (such as pumps, valves, pipelines, etc.)
used in the production process (e.g. Allen et al. 2013).
Clearly, there is a need to constrain the atmospheric flux of
fugitive point and diffuse sources of GHG from gas fields,
not only from an environmental, but also an economic
perspective (e.g. carbon emission pricing mechanisms and
more effective gas recovery).
Developments in cavity ring down spectrometers
(CRDS) have enabled high-resolution, high-precision
measurement of trace gas concentrations and stable
isotope ratios in a field portable device (Kerstel and
Gianfrani 2008; Jackson et al. 2014). This instrumentation has the potential to enable high temporal and spatial
measurements for mapping gas emission hotspots and
to discriminate between potential sources through the
use of stable isotope ratios (Jackson et al. 2014). The
stable isotope ratio of atmospheric gases has been used
to trace local (Krevor et al. 2010), regional (TownsendSmall et al. 2012) and global (Ciais et al. 1995) sources
and sinks of GHGs. Traditionally, stable isotope measurements were measured on discrete samples
using large, expensive, laboratory-restricted isotope ratio mass spectrometers (IRMS). As such,
high temporal and spatial resolution sampling has
been prohibitively laborious, and due to storage
and transport requirements, contamination becomes a potential issue.

Water Air Soil Pollut (2014) 225:2216

In this paper, we describe high-resolution observations of methane (CH4) and carbon dioxide (CO2) within and outside of a production and an exploration CSG
field in Australia. The approach utilised a commercially
available CRDS (Picarro G2201-i) to measure concentrations and carbon stable isotope values (13C) of CH4
and CO2 at high frequency. The results show enrichment
in CH4 and CO2 concentrations in the production field
and localised increases in CH4 and CO2 concentrations
elsewhere. Isotope values indicate a different source of
CH4 and CO2 within the production field as compared to
outside. To our knowledge, our observations represent
the first assessment of GHGs in Australian CSG fields.

2 Material and Methods


Study Area Two separate surveys were conducted over
a week in an exploration field (Casino, New South
Wales, Australia ~28 50 0 S, 153 0 0 E) and within
a CSG production field (Tara, Queensland Australia,
~26 50 0 S, 150 20 0 E). During the Casino survey,
a total of ~300 km was surveyed on 10 August 2012
from 3:41 a.m. to 9:38 a.m. During the Tara survey, a
total of ~700 km was surveyed (~250 km within the
field and ~500 km outside) on 17 August 2012 from
3:31 a.m. to 3:47 p.m. Production and exploration of
CSG in both survey areas are associated with the extensive Walloon Coal Measures within the Surat Basin
(Tara area) and Clarence Moreton Basin (Casino area).
The Walloon Coal Measures are relatively permeable
(>500 mD in some seams) (Scott et al. 2004) and
relatively shallow (near surface to <400 m) (Scott et al.
2007). A total of >300 production wells (see crosses in
Fig. 1), hundreds of kilometres of pipelines and three
compression stations were present in the Tara study site.
In contrast, <50 exploration wells and no associated
infrastructure were present in the Casino region.
CH4, CO2, 13C-CH4 and 13C-CO2
Measurements Measurements during both surveys were
undertaken using a Picarro G2201-i CRDS.
Polyethylene-lined tubing (1/8 in. ID, 1/4 in. OD BevA-Line IV) was attached to a pole protruding ~1.5 m in
front of the vehicle at a height of ~1.5 m. The vehicle
was driven at ~30100 km h1 during the surveys. Since
the instrument provides real-time results, we slowed the
vehicle down near areas of high interest to obtain
higher-resolution data. Sample gas was continuously

Water Air Soil Pollut (2014) 225:2216

Page 3 of 9, 2216

Fig. 1 Spatial survey in the Tara gas field. a CH4 concentrations.


b CO2 concentrations. c Keeling plot of CH4 within the gas field. d
Keeling plot of CH4 outside of the gas field. e Keeling plot of CO2
within the gas field. f Keeling plot of CO2 outside of the gas field.

The crosses indicate the location of CSG wells, and the stars
indicate the location of townships. Note that for clarity, only the
within gas field data are presented and the survey area defined in
the inset of a covers both the Casino and Tara surveys

pumped into the Picarro G2201-i at a rate of


~30 mL min1 by an integrated vacuum pump supplied
with the instrument. The response time of the instrument
to fluctuations of gas concentration at the tubing inlet
was <20 s. Although the instrument does measure water
vapour and correct concentrations accordingly, a column containing desiccant (Drierite) was placed in line
to remove water vapour from the sample stream to
minimise any potential artefacts.
The Picarro G2201-i uses continuous wave cavity
ring down spectroscopy (cw-CRDS) to determine gas
concentrations and isotope ratios. Briefly, a gas sample
stream is continuously pumped through a pressure

(0.19470.0002 atm) and temperature (400.005 C)controlled high-finesse cavity, which contains three
highly reflective mirrors. Light is emitted into the cavity
by a continuous wave laser until a threshold is reached;
then, the laser is switched off. The concentrations of the
individual carbon isotopologues of CH4 and CO2 are
determined by the decay rate (or ring down) of the
isotopologue-specific spectral adsorption line, which is
compared to the cavity only (i.e. without the absorption
gas) ring down rate. The cavity only ring down rate is
continuously measured using a laser tuned to a wavelength where adsorption by sample gas does not occur
(for reviews on infrared isotope spectroscopy, see

2216, Page 4 of 9

Kerstel 2004 and Kerstel and Gianfrani 2008). The


calibration of the instrument was checked prior to and
following each survey using a 502-ppm CO2 standard
(Coregas, Australia), and 13C-CO2 was calibrated by
running the reference gas through both the Picarro
G2201-i and a Delta V+ IRMS. The CH4 standard
had a concentration of 3 ppm (13C-CH4 =45.9),
and the 13C-CH4 value of the standard was obtained
by running the standards through a factory-calibrated
CRDS; it was later referenced against certified standards (Isometric Instruments, Tiso 1, 13C-CH4 =
38.30.2 and Liso1, 13C-CH4 =66.50.2).
No calibration drifts were observed over the timescale
of the experiments (<24 h).
Data Analysis and Mapping Location during each survey was logged at 1 Hz using a handheld GPS (Garmin
GPSMap 76S) interfaced to a notebook computer. The
Picarro G2201-i was also connected to a small notebook
computer via a network cable and remote connection to
enable real-time visual analysis of data. This real-time
data visualisation enabled on the fly decisions to be
made on the course covered during the survey. The
Picarro G2201-i, vacuum pump and small notebook
computers were supplied with power via a bank of four
deep-cycle lead acid batteries (100 amp hour each)
connected in parallel to a 1000 W pure sine wave
inverter (Redarc R12-1000S). This set-up enables
>15 h of continuous operation. The system was run for
at least 7 h prior to the start of each survey to ensure
adequate pressure and temperature stabilisation, although pressure and temperature were generally stable
within 1 h of instrument start-up.
An array of time stamped data from the Picarro
G2201-i was logged at ~1 Hz to an internal hard
drive. The precision of the instrument is increased
with measuring time with a reported 1- precision
for 30-s measurement of 200 ppb (12CO2), 10 ppb
(13CO2) and 50 ppb + 0.05 % for 12CH4. The
reported 13C-CO2 and 13C-CH4 1- precision
values are <0.3 and <0.8, respectively, and for
a 15-min measurement interval, we obtained 30-s
1- precision values of <0.5 and <3.5 for 13C-CO2
and 13C-CH4, respectively. Due to the survey design,
we decided to sacrifice absolute precision for greater
coverage and used a rolling 30-s average for both 13C
and concentrations of CH4 and CO2. The lower absolute
precision of 13C-CO2 and 13C-CH4 values is offset by
the large number of data points (~39,000 for Tara survey

Water Air Soil Pollut (2014) 225:2216

and ~23,000 for Casino survey). Assuming random


noise, the interpretation of general concentration and
isotope relationships (i.e. Keeling plots, see results) is
valid. Data from the Picarro G2201-i was merged with
the data from the GPS, based on the time stamp of the
files. This was then converted to a point shape GIS file
using ArcInfo GIS software.

3 Results and Discussion


Concentrations Both CO2 and CH4 concentrations
were elevated in the Tara CSG field (Fig. 1a, b).
During the Tara survey, background concentrations of
CH4 and CO2 outside the CSG field ranged from ~1.8 to
2 ppm and 390 to 423 ppm, respectively. Inside the gas
field, CH4 concentrations were consistently higher than
2 ppm and reached 6.89 ppm. The gas field CO2 concentrations were consistently higher than 450 ppm and
reached 541 ppm. These values exclude the values from
other obvious potential sources (see below).
During the Casino survey, there were localised increases in both CH4 and CO2 concentrations (up to
2.11 ppm CH4, Fig. 2a, and 457 ppm CO2, Fig. 2b).
The highest concentrations of both CH4 and CO2 were
recorded within 1 km of a CSG well in the southern area
of the survey zone, but it is unclear whether this hotspot
was related to gas leakages or other sources.
Stable Isotopes Keeling plots can be used to determine
the isotope ratio of the added gas, with the intercept of
the regression of the inverse of concentration plotted
against isotope value equal to the average isotope value
of the added gas (Keeling 1958). Keeling plots indicated
that the CH4 source within the CSG field during the Tara
survey (Fig. 1c; 54.700.13) had a distinctly different 13C-CH4 value to the CH4 source outside the CSG
field (Fig. 1d; 47.360.28). The same is true for the
CO2 source within (Fig. 1e; 25.960.06) and outside (Fig. 1f; 31.920.17) the CSG field during the
Tara survey. During the Casino survey, the added CH4
had a 13C value of 56.70.36 and the 13C of the
added CO2 was 28.10.1.
Methane production in the Walloon Coal Measures
(i.e. the target formation of CSG in both the Tara and
Casino gas fields) is predominantly biogenic, driven by
the CO2 reduction pathway (Draper and Boreham
2006). The 13C value of CH4 in regional CSG has a

Water Air Soil Pollut (2014) 225:2216

Page 5 of 9, 2216

Fig. 2 Spatial survey in the Casino region. a CH4 concentrations. b CO2 concentrations. c Keeling plot of CH4. d Keeling plot of CO2. The
crosses indicate the location of CSG wells, and the stars indicate the location of townships

reported narrow range of 57.3 to 54.2 (Draper and


Boreham 2006) which is an intermediate value between
thermogenic methane (generally ~50 to 20 ,
Whiticar 1996) and bacterial-produced methane (~50
to 120, Whiticar 1996, 1999). This value is similar
to the added methane within the Tara gas field (Fig. 1c;
54.700.13) and during the entire Casino survey
(Fig. 1d; 56.70.36). In gas fields where methane is
predominantly thermogenic, CH4 will have a more positive 13C-CH4 value (Whiticar 1996), enabling a greater separation of fugitive emissions from other sources
which generally have a more negative 13C-CH4 value
(Chanton et al. 2005). For example, a recent study that
highlighted fugitive emissions from unconventional gas
fields that may be underestimated by current inventories

(Ptron et al. 2012) has been criticised because it did not


consider the potential influence of agriculture in their
analysis (Sgamma 2012). As demonstrated by our
Keeling plots (Fig. 1), the use of 13C values within
and outside gas fields could help to separate these
sources.
The more enriched 13 C of the added CO 2
within the Tara gas field (25.96 0.06 ;
Fig. 1e) outside the gas field (Fig. 1f; 31.92
0.17) may be driven by a mixture of respiration
of C3 vegetation (the dominant vegetation type in
the area) which would produce CO2 with a 13CCO2 of ~28 to 32 and a source with a more
positive value. While the concentration of CO2 in
the Surat Basin CSG is generally <1 % (Draper

2216, Page 6 of 9

Water Air Soil Pollut (2014) 225:2216

Fig. 3 Concentration of CH4 and CO2 during the surveys when a


passing a vehicle, b driving through a wetland (Tuckean Swamp,
Casino survey), c driving past a combined landfill and sewage

treatment plant site, d driving past a abattoir/ cattle holding facility,


e driving into an urban area (Lismore, NSW Australia, population
~40 000), and f driving through a smoke plume from a bushfire

and Boreham 2006), the average 13C-CO2 of the


CSG is ~19 (Draper and Boreham 2006), and
therefore, fugitive CO2 emissions from the production field may be the source of 13C-enriched CO2

in the Tara area (Fig. 1e). A study in the same


area found higher concentrations of CO2 within
the gas field than a control site outside of the
gas field, which was hypothesised to be linked to

Water Air Soil Pollut (2014) 225:2216

Page 7 of 9, 2216

higher emissions associated with mining-related


changes in soil gas exchange (Tait et al. 2013).
An alternative source of CO2 may be the flux of CO2
from the expansive produced water ponds in the area.
The 13C value of the dissolved inorganic carbon (13CDIC) of produced water from CSG in the Powder River
basin in Wyoming USA was found ranging between +12
and +22 (Sharma and Baggett 2011), with this
enriched value typical of DIC produced through bacterial methanogenesis (Whiticar et al. 1986; Whiticar
1999; Corbett et al. 2012). Assuming ~9 fractionation
between DIC and CO2 (Zeebe and Wolf-Gladrow
2001), the CO2 degassed from these ponds would have
a 13C value of +3 to +13 and therefore could be the
source of 13C-enriched CO2. Further study is required to
quantify the relative importance of this pathway.
It is important to note that the Keeling plot model has
two assumptions: (1) it assumes that the gas of interest in
the atmosphere is simply a function of mixing between
background and a single source, and (2) it assumes that
the isotopic value of these two end members does not
change over the course of observations (Miller and Tans
2003; Pataki et al. 2003). It is likely that at least during
the outside of gas field surveys, these assumptions
were violated due to the following: (1) the large spatial
extent (100 km) and therefore differences in the isotope
value of the source; (2) changes in convective boundary
layer dynamics (surveys were undertaken from early
morning to early afternoon) and therefore differences
in the background concentration and isotope value; and
(3) changes in photosynthesis, respiration and soil gas
fluxes over the survey period as the day progresses. As
the within gas field surveys were undertaken over a
smaller area and under similar boundary layer conditions (i.e. 2 h either side of dawn), it is likely that the
Keeling plot assumptions held true under these conditions. To estimate for differences associated with changes in background isotope and concentration values, we
used the model developed by Miller and Tans (2003);
obs  C obs bg  C bg s C obs C bg

where is the 13C value, C is the concentration, obs is


the observed, bg is the background, and s is the source.
We altered the background concentration and 13C
value to match the lowest concentration and respective
isotope values measured within and outside of the gas

field measured for a period of 5 min and used Model II


regression as recommended by Pataki et al. (2003) and
Miller and Tans (2003) . The results were nearly identical to Keeling plot data with 13C-CH4 source values of
54.94 and 47.59 within and outside of the Tara gas
field, respectively, and 13C-CO2 values of 25.70 and
31.18 within and outside of the Tara gas field,
respectively. We also undertook the same analysis using
a box-car averaging of 30-s data, rather than a running
average. The box-car averaging revealed 13C-CH4
source values of 54.98 and 47.71 and 13C-CO2
values of 25.85 and 31.37 within and outside of
the Tara gas field, respectively.
The consistent values between the three model estimates give us confidence that there is an isotopically
distinct source of CH4 and CO2 within the Tara production gas field and that this source has a signature that is
consistent with the regional CSG signature.
CH4 and CO2 Sources Other than CSG During the two
surveys, a number of other potential point sources were
observed including the following: vehicle emissions
(from passing cars), wetlands, a combined sewage treatment plant and landfill site, an abattoir and cattle holding complex, urban areas, and a bushfire (Fig. 3). Each
of these point sources resulted in an increase of CH4
and/or CO2 concentrations (Fig. 3). With the exception
of the wetland (Tuckean Swamp, Casino survey), the
point sources elicited a similar response for both CH4
and CO2 concentrations. This suggests that the methodology can also be used to map emissions from other
sources of CH4 and CO2.
Implications Much of the difficulty in determining fugitive emissions from gas fields arises from the large
areas that these fields can cover. While measuring leaks
from point sources related to infrastructure (e.g. valves,
well heads, pipes and compressors) may be relatively
straightforward, diffuse gas emissions at the gas field
scale are much more difficult to assess. Changes in
pressure-driven gas transport through existing and
mining-created conduits, related to directional drilling
and hydraulic fracturing, clearly present an additional
challenge in accurately determining emissions from an
entire gas field. Traditional land-air flux methods (e.g.
chambers, gradient boundary layer and eddy covariance) often measure fluxes over a small footprint.
Therefore, methods are required that pinpoint areas of

2216, Page 8 of 9

high fluxes, prior to investing the time and effort into


accurately quantifying fluxes using traditional methods.
By undertaking landscape scale surveys of CO2 and
CH4 concentrations and 13C values using commercially available equipment, areas of potentially high GHG
emissions from unconventional gas fields can be
mapped. While a number of potential CO2 and CH4
sources can exist within a gas field (e.g. when agricultural land uses coexist with gas extraction), the combined approach of using both isotopes and concentrations of CO2 and CH4 can help determine the relative
importance of the various sources.
Data from this
study indicates that unconventional gas may drive
large-scale increases in atmospheric CH4 and CO2 concentrations, which need to be accounted for when determining the net GHG impact of using unconventional gas
sources. Our rapid assessment approach can be used to
design well-targeted experiments to identify and monitor fugitive GHG emissions in areas of high interest and,
with the development of appropriate atmospheric
models, may allow for quantification of emissions at
the landscape scale. Considering the lack of previous
similar studies in Australia, the identified hotspots of
GHGs and the distinct isotopic signature within the Tara
gas field demonstrate the need to fully quantify GHG
emissions before, during and after CSG exploration
commences in individual gas fields.

Acknowledgments The instrumentation used in this study was


funded by the Australian Research Council (LE120100156). DTM
is supported by an S.C.U. Postdoctoral Fellowship.

References
Allen, D. T., Torres, V. M., Thomas, J., Sullivan, D. W., Harrison,
M., Hendler, A., et al. (2013). Measurements of methane
emissions at natural gas production sites in the United
States. Proceedings of the National Academy of Sciences,
110(44), 1776817773. doi:10.1073/pnas.1304880110.
Cathles, L. M., Brown, L., Taam, M., & Hunter, A. (2012). A
commentary on The greenhouse-gas footprint of natural gas
in shale formations by R.W. Howarth, R. Santoro, and
Anthony Ingraffea. Climatic Change 113(525535), doi:10.
1007/s10584-011-0333-0.
Chanton, J. P., Chasar, L. C., Glaser, P., & Siegel, D. (2005).
Carbon and hydrogen isotopic effects in microbial methane
from terrestrial environments. In L. B. Flanagan, J. R.
Ehleringer, & D. E. Pataki (Eds.), Stable isotopes and

Water Air Soil Pollut (2014) 225:2216


biosphere-atmosphere interactions, physiological ecology
series (pp. 85105). Amsterdam: Elsevier-Acadamic Press.
Ciais, P., Tans, P. P., Trolier, M., White, J. W. C., & Francey, R. J.
(1995). A large northern hemisphere terrestrial CO2 sink
indicated by the 13C/12C ratio of atmospheric CO2. Science,
269(5227), 10981102. doi:10.1126/science.269.5227.1098.
Corbett, J. E., Tfaily, M. M., Burdige, D. J., Cooper, W. T., Glaser,
P. H., & Chanton, J. P. (2012). Partitioning pathways of CO2
production in peatlands with stable carbon isotopes.
Biogeochemistry. doi:10.1007/s10533-012-9813-1.
Draper, J. J., & Boreham, C. J. (2006). Geological controls on
exploitable coal seam gas distribution in Queensland. APPEA
Journal, 46, 343366.
Hayhoe, K., Kheshgi, H. S., Jain, A. K., & Wuebbles, D. J. (2002).
Substitution of natural gas for coal: climatic effects of utility
sector emissions. Climatic Change, 54, 107139.
Howarth, R. W., Santoro, R., & Ingraffea, A. (2011). Methane and
greenhouse-gas footprint of natural gas from shale formations. Climatic Change, 106, 679690. doi:10.1007/s10584011-0061-5.
IPCC. (2006). 2006 IPCC guidelines for national greenhouse gas
inventories. Hayama: IPCC.
Jackson, R. B., Down, A., Phillips, N. G., Ackley, R. C., Cook, C.
W., Plata, D. L., et al. (2014). Natural gas pipeline leaks
across Washington, DC. Environmental Science &
Technology, 48(3), 20512058. doi:10.1021/es404474x.
Jaramillo, P., Griffin, W. M., & Mathews, H. S. (2007).
Comparative life-cycle air emissions of coal, domestic natural gas, LNG, and SNG for electricity generation.
Environmental Science & Technology, 41(17), 62906296.
doi:10.1021/es063031o.
Karion, A., Sweeney, C., Ptron, G., Frost, G., Hardesty, R. M.,
Kofler, J., et al. (2013). Methane emissions from airborne
measurements over a western United States natural gas field.
Geophysical Research Letters, 40, 43934397. doi:10.1002/
grl.50811.
Keeling, C. D. (1958). The concentration and isotopic abundances
of atmospheric carbon dioxide in rural areas. Geochimica et
Cosmochimica Acta, 13, 322334.
Kerstel, E. (2004). Isotope ratio infrared spectrometry. In P. A. de
Groot (Ed.), Handbook of stable isotope techniques (pp.
759787). Amsterdam: Elsevier.
Kerstel, E., & Gianfrani, L. (2008). Advances in laser-based
isotope ratio measurements: selected applications. Applied
Physics B: Lasers and Optics, 92, 439449. doi:10.1007/
s00340-008-3128-x.
Krevor, S., Perrin, J.-C., Esposito, A., Rella, C., & Benson, S.
(2010). Rapid detection and characterization of surface CO2
leakage through the real-time measurement of 13C signatures in CO2 flux from the ground. International Journal of
Greenhouse Gas Control, 4, 811815. doi:10.1016/j.ijggc.
2010.05.002.
Miller, J. B., & Tans, P. P. (2003). Calculating isotopic fractionation from atmospheric measurements at various scales.
Tellus, 55B, 207214.
Moore, C. W., Zielinska, B., Ptrone, G., & Jackson, R. B. (2014).
Air impacts of increased natural gas acquisition, processing,
and use: a critical review. Environmental Science &
Technology. doi:10.1021/es4053472.
Osborn, S. G., Vengosh, A., Warner, N. R., & Jackson, R. B.
(2011). Methane contamination of drinking water

Water Air Soil Pollut (2014) 225:2216


accompanying gas-well drilling and hydraulic fracturing.
Proceedings of the National Academy of Sciences, USA,
108(20), 81728176.
Pataki, D. E., Ehleringer, J. R., Flanagan, L. B., Yakir, D.,
Bowling, D. R., Still, C. J., et al. (2003). The application
and interpretation of Keeling plots in terrestrial carbon cycle
research. Global Biogeochemical Cycles, 17(1), 1022. doi:
10.1029/2001GB001850.
Ptron, G., Frost, G., Miller, B. R., Hirsch, A. I., Montzka, S. A.,
Karion, A., et al. (2012). Hydrocarbon emissions characterization in the Colorado Front Range: a pilot study. Journal of
Geophysical Research, 117, D04304. doi:10.1029/
2011JD016360.
Scott, S., Anderson, B., Crosdale, P., Dingwall, J., & Leblang, G.
(2004). Revised geology and coal seam gas characteristics of
the Walloon SubgroupSurat Basin, Queensland. Paper presented at the Eastern Australasian Basins Symposium II.
Scott, S., Anderson, B., Crosdale, P., Dingwall, J., & Leblang, G.
(2007). Coal petrology and coal seam gas contents of the
Walloon SubgroupSurat Basin, Queensland, Australia.
International Journal of Coal Geology, 70, 209222. doi:
10.1016/j.coal.2006.04.010.
Sgamma, K. M. (2012). Colarado methane study not clear-cut.
Nature, 483, 407.
Sharma, S., & Baggett, J. K. (2011). Application of carbon isotopes to detect seepage out of coalbed natural gas produced

Page 9 of 9, 2216
water impoundments. Applied Geochemistry, 26, 1423
1432. doi:10.1016/j.apgeochem.2011.05.015.
Tait, D. R., Santos, I. R., Maher, D. T., Cyronak, T. J., & Davis, R.
J. (2013). Enrichment of radon and carbon dioxide in the
open atmosphere of an Australian coal seam gas field.
Environmental Science & Technology, 47, 30993104. doi:
10.1021/es304538g.
Townsend-Small, A., Tyler, S. C., Pataki, D. E., Xu, X., &
Christensen, L. E. (2012). Isotopic measurements of
atmospheric methane in Los Angeles, California,
USA: influence of fugitive fossil fuel emissions.
Journal of Geophysical Research, 117, D07308. doi:
10.1029/2011JD016826.
Whiticar, M. J. (1996). Stable isotope geochemistry of coals,
humic kerogens and related natural gases. International
Journal of Coal Geology, 32, 191215.
Whiticar, M. J. (1999). Carbon and hydrogen isotope systematics
of bacterial formation and oxidation of methane. Chemical
Geology, 161, 291314.
Whiticar, M. J., Faber, E., & Schoell, M. (1986). Biogenic methane formation in marine and freshwater environments: CO2
reduction vs. acetate fermentationisotope evidence.
Geochimica et Cosmochimica Acta, 50, 693709.
Zeebe, R. E., & Wolf-Gladrow, D. (Eds.). (2001). CO2 in seawater: equilibrium, kinetics, isotopes. Amsterdam: Elsivier
Science B.V.

S-ar putea să vă placă și