You are on page 1of 9

Pagination not final (cite DOI) / Pagination provisoire (citer le DOI)

ARTICLE
Carbon dioxide-induced liberation of methane from
laboratory-formed methane hydrates
Can. J. Chem. Downloaded from www.nrcresearchpress.com by SUNY AT STONY BROOK on 06/14/15
For personal use only.

Kristine Horvat and Devinder Mahajan

Abstract: This paper reports a laboratory mimic study that focused on the extraction of methane (CH4) from hydrates coupled
with sequestration of carbon dioxide (CO2) as hydrates, by taking advantage of preferential thermodynamic stability of hydrates
of CO2 over CH4. Five hydrate formation-decomposition runs focused on CH4CO2 exchange, two baselines and three with host
sediments, were performed in a 200 mL high-pressure Jerguson cell tted with two glass windows that allowed visualization of
the time-resolved hydrate phenomenon. The baseline pure hydrates formed from articial seawater (75 mL) under 6400
6600 kPa CH4 or 28003200 kPa CO2 (hydrate forming regime), when the bath temperature was maintained within 46 C and
the gas/liquid volumetric ratio was 1.7:1 in the water-excess systems. The data show that the induction time for hydrate
appearance was largest at 96 h with CH4, while with CO2 the time shortened by a factor of four. However, when the secondary
gas (CO2 or CH4) was injected into the system containing preformed hydrates, the entering gas formed the hydrate phase
instantly (within minutes) and no lag was observed. In a system containing host Ottawa sand (104 g) and articial seawater
(38 mL), the induction period reduced to 24 h. In runs with multiple charges, the extent of hydrate formation reached 44% of the
theoretical value in the water-excess system, whereas the value maximized at 23% in the gas-excess system. The CO2 hydrate formation
in a system that already contained CH4 hydrates was facile and they remained stable, whereas CH4 hydrate formation in a system
consisting of CO2 hydrates as hosts were initially stable, but CH4 gas in hydrates quickly exchanged with free CO2 gas to form more
stable CO2 hydrates. In all ve runs, even though the system was depressurized, left for over a week at room temperature, and ushed
with nitrogen gas in between runs, hydrates exhibited the memory effect, irrespective of the gas used, a result in contradiction with
that reported previously in the literature. The facile CH4CO2 exchange observed under temperature and pressure conditions that
mimic naturally occurring CH4 hydrates show promise to develop a commercial carbon sequestration system.
Key words: sediment hosted hydrates, gas exchange in hydrates, methane hydrate, carbon dioxide hydrate, carbon sequestration.
Rsum : Cet article prsente les rsultats dune tude visant a` reproduire en laboratoire les conditions dextraction du mthane
(CH4) a` partir dhydrates et, simultanment, la squestration du dioxyde de carbone (CO2) sous forme dhydrates, en tirant parti
de la plus grande stabilit thermodynamique des hydrates de CO2 par rapport aux hydrates de CH4. Nous avons men cinq essais
de formation-dcomposition dhydrates centrs sur lchange entre le CH4 et le CO2, dont deux essais de rfrence et trois,
dans des sdiments contenant des hydrates. Les essais ont t raliss dans une chambre Jerguson a` haute pression de 200 ml
munie de deux fentres de verre permettant lobservation du phnomne de transformation des hydrates en fonction du temps.
Les hydrates purs de rfrence se sont forms a` partir de leau de mer articielle (75 ml) sous une pression de CH4 de 6400 a`
6600 kPa ou de CO2 de 2800 a` 3200 kPa (conditions de formation dhydrates) lorsque la temprature du bain tait maintenue
entre 4 C et 6 C et que le rapport volumtrique gaz/liquide tait denviron 1,7 : 1 dans les systmes o leau tait en excs. Les
donnes rvlent que la priode dinduction requise pour lapparition de lhydrate tait la plus longue, soit 96 h, dans le cas du
CH4, tandis quelle tait quatre fois plus courte dans le cas du CO2. Toutefois, lorsque le gaz secondaire (le CO2 ou le CH4) tait
inject dans le systme contenant des hydrates dja` forms, le gaz entrant sest tout de suite transform en hydrate (en quelques
minutes); aucun dlai na t observ. Dans un systme contenant le substrat dans du sable dOttawa (104 g) et de leau de mer
articielle (38 ml), la priode dinduction tait rduite a` 24 h. Dans les essais a` charges multiples, lhydrate sest form dans une
proportion atteignant 44 % de la valeur thorique en systme o leau tait en excs, tandis que la proportion a atteint une valeur
maximale de 23 % en systme o le gaz tait en excs. La formation dhydrates de CO2 dans un systme qui contient dja` des
hydrates de CH4 tait aise, et les hydrates forms sont demeurs stables. linverse, les hydrates de CH4 forms dans le systme
compos pralablement dhydrates de CO2 taient initialement stables, mais ont rapidement vu leur CH4 gazeux schanger
contre le CO2 gazeux libre pour former des hydrates de CO2 plus stables. Dans les cinq essais, mme dans le cas dun systme tait
dpressuris, laiss a` temprature ambiante pendant plus dune semaine et purg avec de lazote gazeux entre les essais, les
hydrates ont manifest l' effet mmoire , quel que soit le gaz employ, un rsultat qui est en contradiction avec les rsultats
publis auparavant. Lchange ais entre le CH4 et le CO2 observ en conditions de pression et de temprature imitant les
conditions naturelles de formation des hydrates de CH4 semble un moyen prometteur de raliser un systme commercial de
squestration du carbone. [Traduit par la Rdaction]
Mots-cls : hydrates dorigine sdimentaire, changes gazeux dans les hydrates, hydrate de mthane, hydrate de dioxyde de
carbone, squestration du carbone.

Received 7 December 2014. Accepted 15 March 2015.


K. Horvat. Materials Science and Engineering, Stony Brook University, Stony Brook, NY 11794, USA.
D. Mahajan. Sustainable Energy Technologies Department, Brookhaven National Laboratory, Upton, NY 11973, USA.
Corresponding author: Devinder Mahajan (e-mail: dmahajan@bnl.gov).
This article is part of a Special Issue in honour of Dr. John Ripmeester and his outstanding contributions to science.
Can. J. Chem. 93: 19 (2015) dx.doi.org/10.1139/cjc-2014-0562

Published at www.nrcresearchpress.com/cjc on 15 April 2015.

Pagination not final (cite DOI) / Pagination provisoire (citer le DOI)

Can. J. Chem. Downloaded from www.nrcresearchpress.com by SUNY AT STONY BROOK on 06/14/15


For personal use only.

Can. J. Chem. Vol. 93, 2015

Introduction

Materials and methods

With the world population expected to reach over 10 billion by


2050,1 the world is facing an ever increasing need for new energy
sources. Increasing greenhouse gas levels in the atmosphere necessitate the use of carbon neutral energy sources. Current estimates show that over 30 trillion m3 of methane (CH4) is trapped as
hydrates in marine and permafrost environments.2 The presence
of vast CH4 hydrate reserves have been known for decades, but
their extraction, especially from CH4-rich but dispersed marine
hydrates, remains a challenge. Gas production by external stimulation using techniques such as steam injection, gas depressurization, or hydrate-inhibitor injection are now well-established.3 In
marine systems, CH4 harvesting from hydrate may result in
seaoor instability.45 At temperatures below 10 C, carbon dioxide (CO2) hydrates are more thermodynamically stable at lower
pressures than CH4 hydrates,6 and some work has been reported
on the CO2CH4 exchange. Both CH4 and CO2 gases form structure
sI hydrates,7 and the exchange is a fascinating approach by itself:
CH4 is liberated while CO2 is sequestered. However, large-scale
application of the exchange concept requires a thorough understanding of immediate surroundings (sediments) of natural hydrate and other environmental concerns.3
Laboratory studies to test the feasibility of CH4 release from
hydrates by CO2 gas injection are mostly limited to analytical
investigation tools, such as NMR, Raman, and MRI, to provide a
basic understanding of the exchange process.8 Raman spectroscopy established that upon liquid CO2 injection into preformed
CH4 hydrates, there was essentially mol to mol correlation between liberated CH4 and CO2 loss due to CO2 hydrate formation.
In one experiment, 70 mmol CH4 was recovered from hydrate
decomposition, while 71 mmol CO2 was consumed to form hydrates, indicating that CH4 guest molecules were replaced by CO2
in the hydrate structure during the gas exchange.6 Another study9
measured CH4 recovery to be 8.3 mol% in 206 h using gas chromatography (GC). An in situ magnetic resonance imaging (MRI) measurement of a Bentheim sandstone hydrate core sample yielded
50%85% CH4 gas recovery after ushing three times with CO2
using GC.10 Another MRI study of brine solutions in Bentheim
sandstone showed that CH4 hydrate spontaneously converted to
CO2 hydrate when exposed to liquid CO2, and the hydrates were
not found to dissociate to liquid water during the exchange.11
Since CH4 hydrates located in permafrost regions are easier to
access than those in the marine environment, the technology for
carbon sequestration has reached larger-scale eld testing in permafrost regions. Studies have shown that there are an estimated
2.4 trillion m3 of recoverable gas from accessible hydrate accumulations in the North Slope of Alaska alone.12 In 2012, a
ConocoPhilips-led test showed success when a 23 mol% CO2 in
N2 mixture was injected to release over 24 210 m3 of CH4 from a
hydrate reservoir in Alaska. CO2 was preferred over N2 in the
hydrate phase, as about 70% of the N2 gas injected was recovered,
while only 40% of the 1376 m3 of CO2 injected was retrieved.13
Recently, a marine eld test was performed off the coast of California to exchange CO2 in CH4 hydrate. In the experiment, pure
CH4 hydrate was brought to the seaoor from a depth of 690 m
and then enclosed in a cylinder lled with a 25% CO2 75% N2 gas
mixture. The CH4 hydrate was found to dissociate to form a mixed
gas phase, though no CO2 hydrate was found to form under these
simulated conditions.14 Thus, the CH4CO2 exchange reaction
merits further research to develop a feasible method to mine CH4
from natural hydrates. In this paper, we describe several laboratoryscale experiments to understand the CH4CO2 hydrate exchange
kinetics.

Gases
Pressurized cylinders of CH4 and CO2 gases were ordered from
Scott Specialty Gases and Praxair, respectively.
Articial seawater
Articial seawater used to simulate natural medium was prepared according to Kester et al.15
Hydrate former unit description
These experiments were performed in a 200 mL Jerguson cell
tted with two 12 inch long borosilicate windows in the Flexible
Integrated Study of Hydrates (FISH) unit at Brookhaven National
Laboratory (BNL) (Fig. 1) and described in detail elsewhere.16 The
presence of a pressure transducer at the top of the cell and two
thermocouples, located in the gas and liquid phases inside the
cell, allowed for accurate pressuretemperature monitoring during runs. The cell capabilities allowed us to: (i) measure the percentage of hydrates formed in the system, (ii) visualize hydrate
formation and observe their morphology, and (iii) measure composition of free gas above the liquid phase to quantify the mixed
CO2CH4 system. We completed ve runs, as noted in Table 1, in
the Jerguson cell without stirring the cell. The rst two experiments used only 75 mL of articial seawater while the last three
used 38 mL articial seawater and 104 g of 110 m Ottawa sand,17
which was 99.0%99.9% silicon dioxide.18 The cell was pressurized
with gas owing through the bottom such that that the entering
gas bubbled through the articial seawater. The cell was initially
charged with either CO2 or CH4 gas, and after the hydrates started
to form, the second gas was injected into the cell. For example, by
initially forming CH4 hydrates, and then injecting CO2 gas, the
CH4CO2 exchange in hydrates could be monitored. In experiments with preformed CH4 hydrate, the pressure in the cell was
reduced by depressurizing the cell below 4000 kPa prior to CO2
injection to insure that CO2 would be injected as a gas rather than
a liquid. The phase diagram for CO2 and CH4 hydrates in articial
seawater (Fig. 2) served as a guide to follow theoretical monitoring
of the experimental system. We also conducted two runs that
involved the formation of CO2 hydrate followed by a charge with
CH4 gas to allow monitoring the stability of CO2 hydrates if they
were to come into contact with CH4 gas, as could potentially occur
in a free CH4 gas zone or if CO2 hydrates formed near a CH4 plume.
During the depressurization cycle, the system pressure was lowered in multisteps, where in each step the pressure was reduced
several hundred kPa and the system was then allowed to stabilize
for several minutes to an hour prior to subsequent pressure reduction. This process was repeated until the system was totally
depressurized.
Modes of unit operation
The known mole stoichiometry of water/gas in CH4 hydrate and
CO2 hydrate are 5.75:1 and 6:1 respectively. The unit was operated
in either water-excess or gas-excess mode, depending on the mole
ratio of water/gas added to the system. These data were used to
calculate theoretical maximum hydrate from which the extent of
hydrate saturation was calculated by the gas consumed during a
given run.
Gas analysis
GC samples were taken periodically during runs and analyzed
for CH4 and CO2 on a Gow Mac series 580 gas chromatograph
tted with a Supelco Carbonex 1000 45/60 (1.5 m 0.32 mm)
packed column using helium as carrier gas.
Published by NRC Research Press

Pagination not final (cite DOI) / Pagination provisoire (citer le DOI)


Horvat and Mahajan

Can. J. Chem. Downloaded from www.nrcresearchpress.com by SUNY AT STONY BROOK on 06/14/15


For personal use only.

Fig. 1. Schematic of the FISH Unit.

Table 1. Operating conditions for the CH4CO2 exchange runs in a 200 mL Jerguson cell.
Gas

Host media

Water source

PI at
Tbath (kPa)

Tbath
(C)

CH4, followed by CO2


CO2, followed by CH4
CH4, followed by CO2
CO2, followed by CH4
CH4, followed by CO2

None
None
Ottawa sand
Ottawa sand
Ottawa sand

Articial seawater
Articial seawater
Articial seawater
Articial seawater
Articial seawater

6431
2845
6617
3211
6651

4
4
4
6
3

Fig. 2. Phase diagrams of CO2 and CH4 hydrates in articial


seawater (created using the CSMHYD program, Colorado School of
Mines19).

Results and discussion


Run 1: Baseline run with no host, CH4 followed by
CO2 charge
Freshly prepared articial seawater (75 mL) was syringed into
the Jerguson cell, and the cell was pressurized to 7065 kPa with
CH4 gas at room temperature and then cooled till the bath temperature was about 4 C. After several days, an observed pressure
drop of over 2000 kPa was attributed solely to hydrate formation.

After repressurization to 6114 kPa with CH4, the pressure dropped


to 5252 kPa. Together, total gas consumed corresponded to 44% of
the theoretical hydrate value in this excess-water system after two
gas charges. At this point, the cell was switched to the CH4CO2
exchange mode by depressurization to 2866 kPa CH4 and then
pressurized with CO2 gas three times until the cell pressure increased to 4652 kPa, resulting in instant gas hydrate formation.
This was necessitated by the fact that at partial pressures above
4000 kPa, CO2 is in liquid phase. GC samples taken over 16 h
showed >99% CH4 (<1% CO2) in the gas phase that remained essentially unchanged over the next 16 h. Over the next 2 h, the formed
gas hydrates were dissociated by multistep depressurization. The
gas phase CO2 increased sharply as hydrates dissociated, until it
maximized at 95%. The data are consistent with instantaneous
CO2 hydrate formation upon CO2 charges, indicating that the
initial CH4 formation facilitated CO2 hydrate formation, possibly
due to memory effect.20 The hydrate memory effect states that
hydrate induction times are reduced when hydrates are formed
using water in which gas hydrates have previously dissociated.
While there is no discussion of the hydrate memory effect on gas
hydrate formation from two different gases, it is likely that the
presence of CH4 hydrate cages facilitated fast CO2 hydrate formation. Figure 3 shows real-time images of formed mixed gas hydrates.
Run 2: Baseline run with no host, CO2 followed by
CH4 charge
In this run, the articial seawater from the previous run was
used. The cell was ushed with N2 gas and pressurized to 2845 kPa
with CO2 after the bath cooled to 4.5 C. After 22 h, no hydrates
Published by NRC Research Press

Pagination not final (cite DOI) / Pagination provisoire (citer le DOI)

Can. J. Chem. Downloaded from www.nrcresearchpress.com by SUNY AT STONY BROOK on 06/14/15


For personal use only.

Can. J. Chem. Vol. 93, 2015

Fig. 3. Image of mixed CH4CO2 hydrates captured during Run 1.


The images were taken at 312 h when the cell pressure was 4135 kPa
at 3.2 C. The corresponding gas phase composition in the cell was
99% CH4:1% CO2. Note that that mixed gas hydrates have lled the
entire cell viewing region.

Fig. 4. Image taken at 52.8 h, at cell pressure of 6038 kPa at 4 C.


The corresponding gas phase composition was 41% CH4:59% CO2. Gas
hydrates can be clearly seen. The spherical hydrates in the lower
portion of the image formed instantly when CH4 was added to the
cell.

were seen in the cell, and the cell was charged with additional
CO2 gas. Upon repressurization to 3997 kPa at 5.2 C, liquid CO2 was
observed above the aqueous layer in the cell. After two hours,
transparent needle-like CO2 hydrates were observed at the bottom
of the cell, and after an additional hour CO2 hydrates were seen at
the aqueousliquid CO2 meniscus. These hydrates appeared more
like solid ice, unlike the needle-like hydrates still at the bottom of
the reactor. One hour later, the gas was discharged from the cell to
reduce cell pressure to a point where liquid CO2 was no longer
stable, and many of the hydrates present dissociated. The cell was
again repressurized to 3273 kPa with CO2 gas that resulted in
instant hydrate formation during charging. No further pressure
drop was noted after the cell was sealed. Shortly thereafter, the
cell was operated in the CO2CH4 exchange mode by rst depressurization to 1301 kPa and repressurized to 6265 kPa with CH4,
wherein instantaneous CH4 hydrate formation was observed during CH4 charging. Figure 4 shows recorded images of mixed CO2
CH4 hydrates, and the hydrates that formed instantly after the
CH4 injection are clearly seen as small spheres beneath the gas
liquid interface. Though the physical appearance of hydrates of
CH4 and CO2 is indistinguishable, the time resolved analysis of
free gas above the aqueous phase by GC was used to quantify
hydrates in the mixed gas system. Gas samples taken from the cell
one hour after the CH4 charge established the CH4/CO2 ratio to be
35%:65% that changed to 50%:50% over the next 26 h. At this point,
a stepwise depressurization of the system was initiated. The GC
analysis indicated that initially CO2 dominated the gas phase, but
as the cell slowly depressurized, the amount of CH4 in the gas
phase increased. The stability of CH4 hydrates was noted, even
when the partial pressure of CH4 was below the hydrate equilibrium curve (Fig. 2), indicating that complete CH4 hydrate decomposition may be a slow phenomenon.

Run 3: CH4 followed by CO2 charge hosted in Ottawa sand


Run 3 was conducted with Ottawa sand (104 g) as the host that
was fully saturated with articial seawater (38 mL) and to ll space
above the sand pack. After cooling to 4 C, the cell was pressurized
with 6617 kPa CH4 gas, when hydrates were observed at the gas
liquid interface in about 24 h. Over time, the hydrate grew downwards into the solution until they reached the top of the sand pack
(Fig. 5). The pressure drop corresponded to 11% conversion in 48 h
in this excess-gas system. At this point, the cell was depressurized
below the hydrate equilibrium conditions to allow complete hydrate dissociation. The cell was repressurized to 6596 kPa and
within 10 min hydrates were observed as a thin lm on part of the
glass window and at the gasliquid interface. The pressure decreased continuously as hydrates grew from the gasliquid interface downwards into the solution, until the pressure stabilized at
5810 kPa, which corresponded to 12% gas conversion into hydrates
over 2 days. The observed relatively fast hydrate formation is attributed to the memory effect.21
Similar phenomenon of instantaneous hydrate formation was
observed during subsequent CO2 charges in the partial depressurization/repressurization cycle. It is notable that in these runs, the
hydrate formed above the sand pack; few if any, hydrates formed
in the sand pack most likely due to capillary inhibition.22 GC
analysis indicated pure CH4 for several hours but 3 days after the
CO2 injection, the gas phase CH4/CO2 ratio was 86.5%:13.5%. Subsequently, hydrate dissociation was induced by depressurization.
A nal GC sample was taken at 1287 kPa pressure and 3.7 C (below
the hydrate stability zone) that established the gasphase CH4/CO2
ratio of 15%:85%, suggesting that the cell was mostly composed of
CO2 hydrates prior to dissociation.
Published by NRC Research Press

Pagination not final (cite DOI) / Pagination provisoire (citer le DOI)


Horvat and Mahajan

Can. J. Chem. Downloaded from www.nrcresearchpress.com by SUNY AT STONY BROOK on 06/14/15


For personal use only.

Fig. 5. Time-resolved images of hydrate formation during Run 3. (1) Taken 20.1 h after the pre-cooled cell was pressurized with 6382 kPa CH4
at 3.5 C. Prior to (1), the cell was charged with 6617 kPa with CH4 gas. (2) At 194.8 h under 2728 kPa at 3.6 C. Prior to (2), after CO2 hydrates
formed for several days, the cell was partially depressurized, left cooling for 3 days, and then partially depressurized and repressurized several
times. (3) At 236 h under 2542 kPa at 3.6 C. Prior to (3), the CO2 gas was charged twice. Shortly after (3), the cell was depressurized in steps to
bring about complete dissociation of hydrates.

Run 4: CO2 followed by CH4 charge hosted in Ottawa sand


In this run, the Ottawa sand/articial seawater from previous
run were reused after ushing with N2 gas, except that the addition of the two gases was reversed to establish the effect of added
CH4 gas to preformed CO2 hydrates. After pre-cooling to 6 C, the cell
was pressurized to 3211 kPa with CO2, when transparent, needle-like
CO2 hydrates were observed on top of the sand pack in less than 7 h.
After 24 h, the hydrates had spread to the gasliquid interface, and in
the next 24 h hydrates were visible within the host sand. The cell
pressure stabilized and corresponded to 11% conversion of CO2 into
hydrates in this excess-water system. Another CO2 charge resulted in
instantaneous CO2 hydrate formation. Shortly thereafter, CH4 was
injected to increase the cell pressure to 5314 kPa at 3.1 C to observe
the CO2CH4 exchange. The GC analysis established the CH4/CO2
ratio as follows: 6.4%:93.6% (5 min); 30.4%:69.6% (36 min). These data

are consistent with the liberation of CH4 from the hydrate phase
shortly after CO2 was introduced in to the system and displaced CH4
to form CO2 hydrates. A second charge of CH4 increased the cell
pressure to 6134 kPa and the measured CH4/CO2 ratio was 34.9%:
65.1% (5 min) that increased to 47.9%:52.1% over 3 h. It is known that,
initially, CH4 molecules ll both the small and large hydrate cages,
but as time elapses the majority of large hydrate cages are lled with
CO2 that results in increased CH4 concentration in the gas phase.23 A
step-wise depressurization of the cell to bring about hydrate dissociation initially resulted in an increase of CH4 in the gas phase, but as
the pressure decreased the gas phase became richer in CO2. When
the pressure was reduced to 2859 kPa, the gas phase composition was
53.7% CH4:46.3% CO2. However, as the pressure further decreased to
812 kPa, the gas phase became richer in CO2 (65.7%), as the CO2
hydrates began to dissociate. Figure 6 is an image of mixed gas hydrates seen above and in the sand pack.
Published by NRC Research Press

Pagination not final (cite DOI) / Pagination provisoire (citer le DOI)


6

Can. J. Chem. Vol. 93, 2015

Can. J. Chem. Downloaded from www.nrcresearchpress.com by SUNY AT STONY BROOK on 06/14/15


For personal use only.

Fig. 6. Images of hydrate formation taken at 30.4 h under 6010 kPa at 3.1 C. The corresponding gas phase composition was 48% CH4:52% CO2.
Mixed hydrates lled much of the reactor above the same pack though few hydrates were seen in the sand pack.

Run 5: CH4 followed by CO2 charge hosted in Ottawa sand


In Run 5, the Ottawa sand and articial seawater from Run 4
were re-used to establish the exchange of CO2 gas with preformed
hydrates of CH4. The cell was ushed with N2 gas, pre-cooled to
3 C, and then pressurized to 6651 kPa with CH4 gas. Within minutes, a thin coating of ice-like hydrates was seen on the cell glass
windows above the gasliquid interface that spread to the gas
liquid interface after 1 h and then moved downwards into the
solution. The noted pressure drop corresponded to 6% hydrate
saturation in this excess-gas system. After 26 h, the second charge
with CH4 corresponded to an additional 2% gas conversion of CH4
into hydrates. The cell was partially depressurized, then repressurized with CO2 up to 4149 kPa, and then 1 h later, the cell was
again partially depressurized and charged with CO2. Multiple GC
samples over 3 h established the presence of a pure CH4 gas phase.
Upon quick depressurization to 1791 kPa (below the hydrate stability zone), GC analysis established the gas phase rich in CO2
(69.1%), indicating the presence of CO2 trapped as hydrate became
unstable during depressurization. Further depressurization to
405 kPa increased CO2 (81.8%) and decreased CH4 (18.2%) in the gas
phase. Figure 7 shows images of the hydrate growth throughout
run 5.
Comparison of data from runs
This study consisted of 5 experimental runs to understand a
CH4CO2 system in which gas phase CO2 is pumped into a natural

CH4 hydrate reservoir that results in sequestered CO2 as hydrate


with concomitant liberation of CH4. Specically, the completed
5 runs focused on understanding the extent and rates of the CH4
CO2 exchange phenomenon. The exact conditions of all experiments are listed in Tables 1 and 2. Runs 1 and 2 were baseline
(without host sediments), while runs 35 included Ottawa sand as
a host. In runs 1, 3, and 5, the cell was initially charged with CH4,
and once CH4 hydrate formed, CO2 was added to the cell. The
opposite sequence was repeated in runs 2 and 4, wherein CO2 gas
was initially added to the cell, and once CO2 hydrates were observed, CH4 was added to the system.
We extracted induction time data for hydrate formation (the
time when these could be seen by the naked eye) from completed
runs to better understand its dependence on the nature of hydrate
forming gas (CH4 and CO2), host sediment, and memory effect.
The data in Table 2 show that the induction time for CH4 hydrate
appearance was the largest at 96 h (Run 1), while with CO2 the
time shortened to 24 h (Run 2). Runs 1, 3, and 5 followed the same
sequence; initially, CH4 gas was added to the system, and once
hydrates were visible in the reactor, CO2 gas was added. CH4 hydrates formed much more quickly (24 h) in the presence of Ottawa
sand (Run 3) than without it (4 days for Run 1). Similarly, Run 4, in
which CO2 hydrates were formed and then CH4 gas was injected
within the host sand, had a shorter hydrate induction time (7 h)
than Run 2 (24 h), in which no porous media was present. The
Published by NRC Research Press

Pagination not final (cite DOI) / Pagination provisoire (citer le DOI)


Horvat and Mahajan

Can. J. Chem. Downloaded from www.nrcresearchpress.com by SUNY AT STONY BROOK on 06/14/15


For personal use only.

Fig. 7. Time-resolved images of hydrate formation during Run 5. (1) Taken 25 h after the pre-cooled cell was initially pressurized with
6382 kPa CH4 at 3.5 C. Prior to (1), the reactor was charged with 6651 kPa with CH4 gas. After CH4 hydrates formed in 26 h, the cell was
repressurized with CH4 gas to 6686 kPa. (2) At 46.7 h under 6548 kPa of CH4 at 2.9 C. After this image was taken, the cell was partially
depressurized and repressurized with CO2 twice. (3). Taken at 52.2 h under 3797 kPa at 3.0 C. Shortly after (3), the cell was depressurized
quickly to induce hydrate dissociation.

presence of porous media is known to affect the stability of gas


hydrates. Higher pressures and (or) lower temperatures are
needed to form hydrates in sand systems.22 In one instance,
Handa and Stupin24 found that systems containing CH4 or pro-

pane hydrates in silica gel required equilibrium pressures 20%


100% higher than systems without sediments to form hydrates.
The lack of hydrate formation within the Ottawa sand pack herein
is in agreement with these studies. The shortened induction time
Published by NRC Research Press

Pagination not final (cite DOI) / Pagination provisoire (citer le DOI)


8

Can. J. Chem. Vol. 93, 2015

Table 2. Induction time measured when hydrates were rst sighted in


the Jerguson cell.

Can. J. Chem. Downloaded from www.nrcresearchpress.com by SUNY AT STONY BROOK on 06/14/15


For personal use only.

Induction time (h)


Run #

Host
sediments

Gas addition
sequence

Charge 1

Charge 2

1
2
3
4
5

No
No
Ottawa sand
Ottawa sand
Ottawa sand

CH4/CO2
CO2/CH4
CH4/CO2
CO2/CH4
CH4/CO2

96
22a
24
7
<5 min

Instant
2
<0.5
Instant
Instant

Note: The bath temperature was maintained at 36 C.


aNo hydrates were observed prior to CO gas injection during charge 2.
2

observed in runs with Ottawa sand are likely due to sand particles
acting as the nucleation sites for hydrate formation. It is likely
that sand particles became stuck to the walls near the gasliquid
interface in the cell during gas bubbling through the sand pack
during charges, and these impurities affected hydrate nucleation
and induction times.25
A comparison of pure hydrates of CH4 and CO2 in studies found
that CO2 hydrates had shorter induction times than CH4 hydrates.2627 The gas hydrate phase diagram (Fig. 1) shows that in
articial seawater, CO2 hydrates are stable at higher temperatures
and lower pressures than CH4 hydrates. In agreement with previous studies2627 and thermodynamics, the induction times in
Runs 2 and 4, wherein CO2 hydrates were initially formed, were
shorter than the induction times for CH4 hydrates (Runs 1 and 3).
Interestingly, the shortest induction time was noted in Run 5,
wherein CH4 hydrates formed nearly instantly after the cell was
pressurized. This speedy formation is likely due to the hydrate
memory effect. Studies21,28 have shown that hydrate-forming solutions can maintain a memory of their hydrate structure when
warmed up slightly above the hydrate stability region.29 There are
two theories why this occurs: (i) a hydrate frame assembly remains
intact in the solution, either as a partial or ordered conguration
once hydrates had formed and (ii) some gas is left dissolved in the
solution after dissociation.29 For Runs 35 in which the same
Ottawa sand and articial seawater were utilized for hydrate formation, the induction time was signicantly reduced even though
different gases (CH4 or CO2) were initially used to form hydrates.
In Run 3, CH4 hydrates formed after about 24 h, while in Run 4,
CO2 hydrates formed in 7 h, and in Run 5, CH4 hydrates formed
within minutes. In between each of these runs, the cell was completely depressurized and ushed with nitrogen gas. Mazloum
et al.30 found that the depressurization of a system of natural gas
hydrates, followed by repressurization with fresh natural gas, did
not result in the hydrate memory effect for gas hydrate formation.
It was found that the hydrate memory effect was destroyed by
depressurization to atmospheric pressure, but reducing the pressure of the system even slightly would still result in the hydrate
memory effect.30 In addition, if the solution is warmed above 2521
or 28C29 or for too long a time period (several hours29), this
retention effect will not occur. In between each of these experiments the cell was left uncooled for extended periods of time
ranging from 7 days to 2 months. All of these factors are contrary
to the decreased hydrate formation time observed herein with the
systems composed of CH4 and CO2.
The sequence of gas addition to the cell affected hydrate stability. In all experiments, once gas hydrate had formed, any additional gas charge, independent of the gas (CO2 or CH4), resulted in
near instant hydrate formation. This fast formation indicates that
if CO2 is to be sequestered in a CH4 hydrate reservoir, it is likely
that CO2 hydrates will form instantly. In addition, whether CO2 or
CH4 was the initial gas for hydrate formation, when a second gas
was charged into the reactor, initially the gas phase of the cell was
mostly, if not entirely, composed of the initial system gas, while
the newly injected gas entered the hydrate phase. For systems

where the cell was initially lled with CH4 and then charged with
CO2 gas, the gas phase of the cell composed of pure CH4 for several
hours after the CO2 injection. Whereas for systems originally pressurized with CO2 and then injected with CH4, the gas phase was
initially mostly composed of CO2, but over a much shorter time
period, the percentage of CH4 in the gas phase increased, indicating that CO2 gas exchanged with CH4 gas in the hydrate structure.
Uchida et al.23 reported similar results with the following explanation: during initial stages of the hydrate formation, CH4 molecules are able to occupy both small and large hydrate cages, but
over time the larger cages mostly trap CO2 gas, resulting in an
increased CH4 in the gas phase.23 The phase eld simulation models have yielded similar results.31 Overall, whether CH4 or CO2 gas
is introduced into a system of gas hydrates via bubbling, an instant formation of hydrates of injected gas is likely to occur.

Conclusions
This study examined preliminary characteristics of the CH4
CO2 exchange phenomenon. Several noteworthy observations are
as follows: (i) the induction times of hydrate formation were generally shorter for CO2 than CH4, by as much as by a factor of four,
with or without sediments; (ii) the hydrate formation data show
that when a secondary gas was injected into a system containing
preformed hydrates, the entering gas formed the hydrate phase
instantly (within minutes); (iii) CO2 hydrates formed in a system
that already contained CH4 hydrates were found to be more
stable, whereas CH4 hydrates formed in a system consisting of
CO2 hydrates as hosts were initially stable, but CH4 gas in hydrates
quickly exchanged with free CO2 gas to form more stable
CO2 hydrates; (iv) the hydrate memory effect was noted during the
3 runs performed to form sand-hosted gas hydrates. In all 5 runs,
even though the system was depressurized, left over a week at
room temperature, and ushed with nitrogen gas in between
runs, the system still exhibited the memory effect. These results
contradict those previously reported in the literature. In summary, the observed fast CO2 hydrate formation from free CO2 gas
in the presence of preformed CH4 hydrate indicate the feasibility
of developing a CO2 sequestration scheme using natural CH4 hydrate reservoirs. Our ongoing and planned work includes runs
that will quantify the extent of uptake and liberation of gases
through the CH4CO2 exchange as a function of key hydrate parameters and at the micrometer scale using X-ray computed microtomography.

Acknowledgements
The authors thank the Ofce of Vice-President of Research
(OVPR) at Stony Brook University for providing funds for the work.
The work was partially supported by the Program Development
funds at Brookhaven National Laboratory.

References
(1) Sulston, J.; Rumsby, M.; Green, N. Environ. Resour. Econ. 2013, 55, 469. doi:10.
1007/s10640-013-9681-8.
(2) Boswell, R.; Collett, T. S. Energy Environ. Sci. 2011, 4 (4), 1206. doi:10.1039/
C0EE00203H.
(3) Collett, T.; Bahk, J.-J.; Baker, R.; Boswell, R.; Divins, D.; Frye, M.; Goldberg, D.;
Huseb, J.; Koh, C.; Malone, M.; Morell, M.; Myers, G.; Shipp, C.; Torres, M.
J. Chem. Eng. Data 2015, 60 (2), 319. doi:10.1021/je500604h.
(4) Collett, T.; Bahk, J.-J.; Frye, M.; Goldberg, D.; Husebo, J.; Koh, C.; Malone, M.;
Shipp, C.; Torres, M.; Myers, G.; Divins, D.; Morell, M. 2013; p Medium: ED.
(5) Priest, J. A.; Clayton, C. R. I.; Rees, E. V. L. Mar. Petrol. Geol. 2014, 58, 187.
doi:10.1016/j.marpetgeo.2014.05.008.
(6) Ota, M.; Morohashi, K.; Abe, Y.; Watanabe, M.; Smith, R. L.; Inomata, H.
Energy Convers. Manage. 2005, 46 (1112), 1680.
(7) Ohgaki, K.; Takano, K.; Moritoki, M. Kagaku Kogaku Ronbunshu 1994, 20 (1),
121. doi:10.1252/kakoronbunshu.20.121.
(8) Komatsu, H.; Ota, M.; Smith, R. L., Jr.; Inomata, H. J. Taiwan Inst. Chem. Eng.
2013, 44 (4), 517. doi:10.1016/j.jtice.2013.03.010.
(9) Jadhawar, P.; Mohammadi, A. H.; Yang, J.; Tohidi, B. Subsurface carbon dioxide
storage through clathrate hydrate formation; In Advances in the Geological Storage
Published by NRC Research Press

Pagination not final (cite DOI) / Pagination provisoire (citer le DOI)


Horvat and Mahajan

(10)
(11)

(12)
(13)

Can. J. Chem. Downloaded from www.nrcresearchpress.com by SUNY AT STONY BROOK on 06/14/15


For personal use only.

(14)
(15)
(16)
(17)

(18)
(19)

of Carbon Dioxide. Nato Science Series: IV: Earth and Environmental Sciences, 2006;
Vol. 65, pp. 111126.
Graue, A.; Kvamme, B.; Baldwin, B. A.; Stevens, J.; Howard, J.; Aspenes, E.;
Ersland, G.; Husebo, J.; Zornes, D. SPE J. 2008, 146.
Baldwin, B. A.; Stevens, J.; Howard, J. J.; Graue, A.; Kvamme, B.; Aspenes, E.;
Ersland, G.; Husebo, J.; Zornes, D. R. Magn. Reson. Imaging 2009, 27, 720.
doi:10.1016/j.mri.2008.11.011.
White, R.; Kumagai, T. Platts Gas Daily 2011, 28, 206.
Schoderbek, D.; Farrell, H.; Hester, K.; Howard, J.; Raterman, K.;
Silpngarmlert, S.; Martin, K. L.; Smith, B.; Klein, P. ConocoPhilips Company,
2013.
Brewer, P. G.; Peltzer, E. T.; Walz, P. M.; Coward, E. K.; Stern, L. A.;
Kirby, S. H.; Pinkston, J. Energy Fuels 2014, 28 (11), 7061. doi:10.1021/ef501430h.
Kester, D. R.; Duedall, I. W.; Connors, D. N.; Pytkowic, R. M. Limnol. Oceanogr.
1967, 12 (1), 176. doi:10.4319/lo.1967.12.1.0176.
Horvat, K.; Kerkar, P.; Jones, K.; Mahajan, D. Energies 2012, 5, 2248. doi:10.
3390/en5072248.
Santamarina, J. C.; Ruppel, C. In Proceedings of the 6th International Conference
on Gas Hydrates (ICGH 2008), Vancouver, British Columbia, Canada, July 610, 2008;
Vancouver, British Columbia, Canada, 2008.
U.S. Silica Company. Material Safety Data Sheet Silica Sand and Ground Silica.
Colorado School of Mines. Software; Available from http://hydrates.mines.
edu/CHR/Software.html.

(20) Bishnoi, P. R.; Natarajan, V. Fluid Phase Equilib. 1996, 117 (12), 168. doi:10.
1016/0378-3812(95)02950-8.
(21) Wu, Q.; Zhang, B. J. Nat. Gas Chem. 2010, 19 (4), 446. doi:10.1016/S10039953(09)60086-4.
(22) Clennell, M. B.; Hovland, M.; Booth, J. S.; Henry, P.; Winters, W. J. J. Geophys.
Res.: Solid Earth 1999, 104 (B10), 22985. doi:10.1029/1999JB900175.
(23) Uchida, T.; Ikeda, I. Y.; Takeya, S.; Kamata, Y.; Ohmura, R.; Nagao, J.;
Zatsepina, O. Y.; Buffett, B. A. ChemPhysChem 2005, 6 (4), 646. doi:10.1002/
cphc.200400364.
(24) Handa, Y. P.; Stupin, D. Y. J. Phys. Chem. 1992, 96 (21), 8599. doi:10.1021/
j100200a071.
(25) Jensen, L.; Thomsen, K.; von Solms, N. Chem. Eng. Sci. 2008, 63 (12), 3069.
doi:10.1016/j.ces.2008.03.006.
(26) He, Y.; Rudolph, E. S. J.; Zitha, P. L. J.; Golombok, M. Fuel 2011, 90 (1), 272.
doi:10.1016/j.fuel.2010.09.032.
(27) Golombok, M.; Ineke, E.; Luzardo, J. C. R.; He, Y. Y.; Zitha, P. Environ. Chem.
Lett. 2009, 7 (4), 325. doi:10.1007/s10311-008-0173-y.
(28) Zhang, B. Y.; Wu, Q.; Gao, X.; Liu, C. L.; Ye, Y. G. Int. J. Environ. Pollut. 2013, 53
(34), 201. doi:10.1504/IJEP.2013.059913.
(29) Sloan, E. D.; Koh, C. A. Clathrate hydrates of natural gases, 3rd ed.; CRC Press:
Boca Raton, FL, 2008.
(30) Mazloum, S.; Yang, J.; Chapoy, A.; Tohidi, B. In Proceedings of the 7th International
Conference on Gas Hydrates, Edinburgh, Scothland, United Kingdom, Edinburgh,
Scothland, United Kingdom, 2011.
(31) Qasim, M.; Kvamme, B.; Baig, K. Int. J. Geol. 2011, 5 (2), 48.

Published by NRC Research Press

You might also like