Вы находитесь на странице: 1из 8

Carbon 42 (2004) 12491256

www.elsevier.com/locate/carbon

Improved methane storage capacities by sorption


on wet active carbons
A. Perrin, A. Celzard *, J.F. Mar^eche, G. Furdin
Laboratoire de Chimie du Solide Mineral, Universite Henri Poincare Nancy I, UMR CNRS 7555, BP 239, 54506 Vandoeuvre-les-Nancy, France
Received 10 July 2003; accepted 28 December 2003
Available online 10 February 2004

Abstract
The possibility of storing large amounts of natural gas within wet active carbons is examined. The sorption isotherms of methane
at 2 C and up to 8 MPa are built for four carbonaceous materials. Three of them originate from the same precursor (coconut shell),
are physically activated at various burn-os and are mainly microporous. The fourth material is a highly mesoporous chemically
activated pinewood carbon. These adsorbents are wetted with a constant weight ratio water/carbon close to 1. The resulting isotherms all exhibit a marked step occurring near the expected formation pressure of methane hydrates, thus supporting their
occurrence within the porous materials. The amount of gas stored at the highest pressures investigated then ranges from 6 to 17 mol/
kg of wet adsorbent (i.e., corresponding to 1036 mol/kg of dry carbon), depending on the material. The results are discussed on the
basis of the known pore texture of each adsorbent, and stoichiometries of the formed hydrates are calculated. Considerations about
adsorption/desorption kinetics and metastability are also developed.
 2004 Elsevier Ltd. All rights reserved.
Keywords: A. Activated carbon; C. Adsorption; D. Gas storage

1. Introduction
Since methane is a cheap and abundant fuel, its
storage and transportation is of great interest [1]. It is
well known that methane can be stored by adsorption in
active carbons, however the commonly accepted target
[2] of 150 STP volumes of gas per volume of storage
vessel (V/V) deliverable at 3.5 MPa and room temperature is still hardly reached. Recently, a new route was
proposed by two Japanese patents [3,4] dealing with
methane sorption on wet active carbons. Within the
porosity of these materials, clathrate hydrates were
suspected to occur, leading to very high storage capacities, close to 300 V/V. Indeed, in such structures, CH4
units are trapped in cages made of water molecules
linked with each other through hydrogen bonding. A
Chinese team rst tried to reproduce these results, using
the same experimental conditions (low amounts of

water, moderate pressures and room temperature), but


only very little eects of wetting the adsorbents were
found [5]. Later, the same authors could nally form
clathrate hydrates within wetted active carbons, but the
experimental conditions which were required were more
realistic than those stated by the Japanese patents. Thus,
methane uptakes close to 200 V/V could be obtained at
10 MPa and 2 C [6]. In the present work, such results
could be reproduced and even improved. The materials,
which were studied both dry and water-wetted, and the
measuring apparatus are described in Section 2, while
storage isotherms are given and discussed in Section 3.
Formation kinetics, stoichiometry and stability of hydrates are discussed in Section 4. Finally, the practical
problems of storing methane through hydrate formation
are examined in the conclusion.

2. Experimental
*

Corresponding author. Tel.: +33-3-83-68-46-30; fax: +33-3-83-6846-19.


E-mail address: alain.celzard@lcsm.uhp-nancy.fr (A. Celzard).
0008-6223/$ - see front matter  2004 Elsevier Ltd. All rights reserved.
doi:10.1016/j.carbon.2004.01.039

The present study deals with a few commercial active


carbons (see Table 1), produced and supplied by the
French company Pica. These materials were ground and

1250

A. Perrin et al. / Carbon 42 (2004) 12491256

Table 1
Pore texture (in the dry state) of the carbonaceous adsorbents
Carbon

SBET

Vl

Vm

Vl Vm

ddry

dwet

NC58
NC86
NC120
Picazine

1000
1587
2031
1967

0.405
0.570
0.814
0.65

0.049
0.088
0.141
0.72

0.454
0.658
0.955
1.37

0.4
0.34
0.29
0.16

1.27
0.98
0.85
0.54

SBET is the BET surface area (m2 /g), Vl and Vm are the micropore and the mesopore volumes (cm3 /g), respectively, while ddry and dwet are the packing
densities (g/cm3 ) of dry and wet adsorbents, respectively.

3. Sorption isotherms
3.1. General comments
At 2 C, wetted carbons exhibit stepwise isotherms,
just like those already built by the Chinese team [6], see
Fig. 1. The methane uptake increases slowly with the
pressure, and then jumps at a critical pressure ranging
typically from 3 to 4 MPa, depending on the material.
Given their characteristic shape, the rst parts of the
curves are attributed to a classical physisorption of
methane within the smallest pores of the adsorbents.
Moreover, below 3 MPa, the data points are accurately
tted by the Freundlich equation which holds for energetically heterogeneous surfaces [7]. From Fig. 1, it may
be seen that the methane uptake is very low below 3
MPa. Indeed, the amounts of methane stored in the
pressure range 03 MPa are much higher for dry carbons, as seen in Fig. 2 and Table 2. The low performances of wetted materials are attributed to the
presence of water spreading in the pore space, hence
making a number of micropores unavailable for the gas.
Wetting the adsorbents is thus unfavourable as far as
moderate pressures are concerned.
However, within the range 34 MPa, the amounts of
methane stored increase rapidly and become higher than

15

Methane uptake (mol/kg wet )

sieved; for reproducibility concerns, only the grains


within the range 100200 lm were handled and investigated. Three active carbons, named NC, are derived from
coconut shells; they were physically activated with steam,
in such a way that the higher the number of the material,
the higher the burn-o. For example, NC86 was more
activated than NC58. The last carbon, called Picazine, is
derived from pinewood, and was chemically activated
with orthophosphoric acid. Before being tested in the
experiments below, Picazine was washed thoroughly with
water using a soxhlet. The main features of the four dry
adsorbents, namely BET surface area, pore texture and
apparent densities, are given in Table 1. As expected, the
NC series is especially microporous, with increasing pore
volumes from NC58 to NC120. Picazine is much more
mesoporous, however the micropore volume is still
important, being almost similar to that of NC86.
The powdery adsorbents were poured into a stainless
steel sample holder having a volume close to 24 cm3 . A
controlled amount of water was then added, in such a
way that the ratio of the water weight to that of the dry
carbon was close to 1. The apparent densities of the
wetted carbons are reported in Table 1. Next, the vessel
was closed and introduced in a cooling device, with
which the temperature of the wetted adsorbent could be
kept at 2 0.1 C.
While investigated in the dry state, the adsorbents
were outgassed at 200 C overnight. However, due to the
presence of controlled amounts of water within the
storage cell, the wetted materials could not be outgassed
at all. The amount of methane stored within each
adsorbent was measured up to 8 MPa using a pressurebearing volumetric apparatus, equipped with a highpressure transducer. The methane uptake was calculated
point to point by discontinuous introduction of the
adsorbate into the sample holder. For that purpose,
the compression of the gas outside of the vessel, i.e., in
the parts of the apparatus free of adsorbent, was taken
into account in the calculations. The number of moles of
methane really stored in the sample holder was obtained
by application of the Van der Waals equation of state of
methane. At the end of each isotherm, the calculated
methane uptake at 8 MPa was compared with the
amount of gas released which could be measured by
opening and weighing the vessel. A very good agreement
was observed for each experiment.

NC58 wet
NC86 wet
NC120 wet
Picazine wet

10

0
0

Peq(MPa)
Fig. 1. Massic methane uptakes of the four carbonaceous adsorbents
listed in Table 1, at 2 C and wetted with a constant weight ratio water/
carbon 1.

A. Perrin et al. / Carbon 42 (2004) 12491256


250

50
NC58 dry
NC86 dry
NC120 dry
Picazine dry

40

Methane uptake (V/V)

Methane uptake (mol/kg dry )

1251

30

20

NC86 wet
NC86 dry
NGC

200

150

100

10

50

0
0

10

P (MPa)
eq

P (MPa)
eq

Fig. 2. Same as Fig. 1, but the four carbonaceous adsorbents are dry.

those measured for dry carbons. As an example, a direct


comparison is given between dry and wetted NC86 in
Fig. 3. Such a step in the storage isotherms of wetted
materials is attributed to the occurrence of clathrate
hydrates. Methane hydrates should be formed rst in
the largest pores, for which the formation pressures are
the lowest. Indeed, the formation pressure at 2 C in
bulk water is Pw 3:14 MPa [8]. Now, water is not free
within wet carbons but trapped in pores which are prone
to capillary eects. Therefore, forming hydrates in these
pores requires an additional capillary pressure DP , and
hence the formation pressure Pf reads
Pf Pw DP :

DP is related to the pore diameter d through the Young


Laplace equation:
DP

4rlh cos h
;
d

where rlh 26:7  103 J/m2 [9,10] is the surface tension


between liquid water and hydrate phases, onto which
water spreads with a contact angle h assumed to be zero.
Consequently, low formation pressures correspond to
large pore sizes and vice-versa [11]. Such a behaviour
was already observed with clathrate hydrates prepared
in silica gels, and for which the formation pressures are
higher than those in free water at the same temperature
[9,12]. Using Eqs. (1) and (2), the pore widths within
which the hydrates rst occur in each adsorbent, i.e., at

Fig. 3. Volumic capacities of the adsorbent NC86, either dry or wetted. The data corresponding to compressed natural gas (CNG) are
given for comparison.

their respective formation pressures, are easily calculated. The results are listed in Table 3. It is found that
hydrates are rst formed in macropores (d > 50 nm),
hence it seems that microporous or even mesoporous
carbons are not the most suitable adsorbents for storing
methane by hydrate formation. Additionally, it can be
seen that, going through the NC series from NC58 to
NC120, increasing pore widths are found, in agreement
with the fact that these materials were increasingly
activated.
Further increasing the pressure induces the formation
of hydrates within pores smaller than those quoted in
Table 3, each pore size having its own formation pressure. At pressures ranging from 5 to 8 MPa, the methane uptakes are either almost constant (for NC58) or
slightly increasing (for the other carbons), see Fig. 1. It
may be supposed that constant uptakes correspond to
Table 3
Formation pressures Pf at 2 C within the pores of wetted materials
having the diameters d calculated using Eqs. (1) and (2)
Carbon

Pf (MPa, 2 C)

d (nm)

NC58
NC86
NC120
Picazine

4.0
3.5
3.49
3.7

124
297
305
191

Table 2
Storage capacities, expressed both in mol/kg of adsorbent and in V/V, of the studied carbonaceous materials, either dry or wetted by water
Carbon

Amount of methane stored at 8 MPa and 2 C


Dry adsorbents

NC58
NC86
NC120
Picazine

Wet adsorbents

mol/kgdry

mol/kgdry ddry

V/V

mol/kgdry

mol/kgwet

mol/kgwet dwet

V/V

12.8
21
23.1
40

5.1
7.1
6.7
6.4

107
158
121
164

10.5
16.5
21.9
35.7

5.8
9.2
10.9
17.4

7.4
9.0
9.2
9.4

177
216
223
227

A. Perrin et al. / Carbon 42 (2004) 12491256

adsorbents in which the pores are completely lled. Indeed, the pores in which the water could not penetrate
(i.e., the smallest ones) are lled with physisorbed
methane, while those in which water was present (i.e.,
the largest ones) are lled by crystallized hydrates. For
both kinds of pores, no additional methane can be
stored in the adsorbent. Conversely, increasing uptakes
at increasing pressures correspond to adsorbents in
which more methane can be stored, essentially through
compression of the gas in formerly empty pores. Indeed,
the smallest pores were already lled by adsorbed
methane in the rst part of the isotherms, while the
largest pores were lled by hydrates above Pf . However,
if the materials were not wetted enough, some pores
which are too large for adsorbing methane are free of
water, and hence can accommodate compressed gas.
Indeed, all the materials were wetted with the same
weight ratio water/carbon 1, irrespective to their pore
volumes. Thus, NC58 was the only adsorbent which
pore space was almost saturated with water. At the
present time, the authors are rebuilding the storage
isotherms of each carbon after saturation with water.
The rst results agree with the above explanations.
3.2. Storage performances
In Figs. 1 and 2, the results were presented in units of
moles stored per weight of adsorbent. However, since
each powdery adsorbent has its own apparent density,
the number of STP volumes of gas stored is not proportional to the mass uptake. The isotherms thus have
to be recalculated, using the densities given in Table 1,
and the results have to be compared with the target of
150 V/V deliverable early suggested by the Atlanta Gas
Light Adsorbent Research Group (AGLARG) for a
viable application to natural gas vehicle [2]. In Fig. 4,
the data of Fig. 1 were expressed in units of V/V. It can
be seen that the isotherms are very similar to each other,
whatever the adsorbent. Such a nding is not surprising
if the apparent densities of the materials are examined:
the more activated the carbon (i.e., the larger its pore
volume), the lower its packing density. This statement
holds for both dry and wetted materials. Now, the volume capacity is proportional to the product mass
capacity packing density, and it can be noted that such
a product, even if not really constant, varies much less
than the mass capacity (see Table 2). Hence, the
adsorbent exhibiting the best performances is that which
combines both the highest apparent density and the
highest mass capacity. Unfortunately, since the adsorbents are the results of an activation process, these two
properties vary in opposite ways. The compensation
between density and mass capacity nally explains why
the volume uptakes seen in Fig. 4 are so similar.
The values of V/V gathered in Table 2 show that the
AGLARG target value of 150 is exceeded with all the

250

Methane uptake (V/V)

1252

NC58 wet
NC86 wet
NC120 wet
Picazine wet

200

150

100

50

0
0

P (MPa)
eq

Fig. 4. Same as Fig. 1, but the methane uptake is expressed in units


of V/V.

wetted materials, by contrast with the dry adsorbents.


However, it should be recalled that high pressures (8
MPa) and low temperatures (2 C) were required, while
the target value was dened at 3.5 MPa and room
temperature. Consequently, it seems that methane
storage by clathrate hydrate formation still not meets
what was recommended by the AGLARG. Nevertheless, hydrates possess one serious advantage in terms of
deliverable amounts. Indeed, while the sorption isotherms of dry materials (see Fig. 2) evidence that a nonnegligible quantity of methane remains adsorbed at 0.1
MPa, those of wetted carbons show that the methane
may be released almost completely.

4. Properties of methane hydrates occurring within wetted


active carbons
4.1. Formation kinetics
In the previous section, the rst parts of stepwise
isotherms were attributed to physisorption, while hydrate formation was invoked to account for the second
parts. Such a description is strongly supported by the
sorption kinetics, which are drastically dierent below
and above Pf . Indeed, introducing a given pressure of
adsorbate into the sample holder, the equilibrium is
usually reached within a few tens of minutes, as far as
the nal equilibrium pressure is lower than Pf . These
adsorption kinetics are very similar to those corresponding to dry materials. Now, if the nal equilibrium
pressure is greater than Pf , the sorption kinetics are
tremendously long, and several days (and sometimes
more than one week) are required. It seems that such
kinetics are typical of hydrate formation, and were already observed within mesoporous silicas [12]. Besides,
the higher the methane uptake (i.e., the higher the

A. Perrin et al. / Carbon 42 (2004) 12491256

lnV =Veq  kt:

V is the STP volume of stored methane measured as a


function of time, Veq is the adsorbed amount at equilibrium, t is the time and k a kinetic constant ranging
from 103 to 104 min1 . It should be noticed that such
values were already observed in other porous media [13]
and thus seem to be very typical. Plotting lnV =Veq as a
function of time indeed evidences straight lines, as seen
in Fig. 5 for several adsorbents.
Once hydrates are formed in the largest pores,
introducing more methane into the sample holder forces
the gas to diuse throughout the pore networkpartly
but more and morelled by hydrate crystallites. This
is the reason why the formation kinetics become
increasingly slower while the nal equilibrium pressure
Peq rises. The following exponential laws (in which k and
Peq are expressed in MPa and in min1 , respectively)
were indeed observed for the NC series and for Picazine,
respectively:
k  6:03  103 exp0:23 Peq
;

4a

k  4:02  104 exp0:06 Peq


:

4b

The relevance of Eqs. (4a) and (4b) is shown in Fig. 6.

6.0
5.5
5.0

Ln(V/Veq )

4.5
4.0
3.5
3.0
2.5
2.0

NC58
NC86
NC120
Picazine

1.5
1.0
0.5
0.0
0

200

400

600

800

1000

1200

1400

1600

Time (min)
Fig. 5. First-order kinetic laws accounting for hydrate formation,
evidenced above Pf for several adsorbents.

0.01

NC
Picazine

-1

k (min )

equilibrium pressure), the slower the sorption kinetics.


This nding is in agreement with the above picture of
pore lling by crystallized hydrates: the latter rst occur
in the largest pores at a pressure Pf , then are formed in
smaller pores at higher pressures. But such smaller
pores, which were formerly accessible through the
largest ones, are now lled by hydrates. Hence, the
greater the equilibrium pressure, the lower the accessibility to smaller pores, and thus the slower the formation kinetics. Moreover, above Pf and whatever the
materials, a rst-order kinetic law was evidenced, which
reads

1253

0.001

0.0001
4

4.5

5.5

6.5

7.5

Peq (MPa)
Fig. 6. Exponential slowing down of the hydrate formation kinetics as
the equilibrium pressure increases.

4.2. Stoichiometry of the hydrate phases


In bulk water, the presence of methane in suitable
conditions (i.e., low temperature and high pressure) is
expected to lead to only one phase, called SI , which has a
clathrate structure and the stoichiometry 8CH4 , 46H2 O
[8]. Within large pores, i.e. macropores and even mesopores, water may be assumed to behave like a liquid and
consequently, the stoichiometry of the SI phase should be
relevant. However, in very narrow pores, it has been
supposed that methane is preferentially adsorbed onto
the carbon walls, and that water molecules come and
build half and/or complete cages around CH4 units,
depending on the pore width [14]. In such conditions,
while the structure remains unchanged, more methane
can be stored by hydrate formation in very small pores,
and nanohydrates having extreme stoichiometries like
CH4 , 2H2 O could be formed [14]. Considering the
heights of the steps of the isotherms shown in Fig. 1, and
knowing the amount of water that was added to each
adsorbent, the stoichiometries of the formed hydrates
may be calculated. However, such a calculation assumes
that the water was completely consumed by hydrate
formation. For the carbon NC58, the calculated composition of the hydrate is 5.8CH4 , 46H2 O. Since the
presence of empty cages is unknown in the SI structure,
whatever the pore size [15], such an average stoichiometry probably means that some water, not accessible to
methane or trapped in very small pores, was not consumed by hydrate formation. The same calculation,
based on the step heights, lead to the following stoichiometries: 8.9CH4 , 46H2 O; 11.4CH4 , 46H2 O; and
18.6CH4 , 46H2 O for NC86, NC120 and Picazine,
respectively. Now, these average stoichiometries comprise more methane than what is expected for the SI
phase. Assuming that all the water is consumed by hydrate formation, such a nding suggests that more
methane can be stored in the adsorbents due to a number of empty pores. Indeed, as already claimed in the

1254

A. Perrin et al. / Carbon 42 (2004) 12491256

above discussion dealing with increasing uptakes above


Pf , the pore spaces were not saturated by water. Consequently, the whole water was accessible to form hydrates,
and gaseous methane could be compressed in the
remaining voids. Thus, even if the average stoichiometry
calculated for Picazine is close to that predicted for the
nanohydrates, i.e., CH4 , 2H2 O [14], a pure hydrate phase
should not be invoked: a mixture of SI phase plus compressed gas is much more convincing. Indeed, coming
back to Eqs. (1) and (2), it may be calculated that, at the
maximum pressure applied (close to 8 MPa), the smallest
pores in which hydrates may be formed are 22 nm
wide. Now, such width is about 20 times larger than what
is required to form the nanohydrates [14].
4.3. Metastability of the hydrate phases
After the building of the isotherms, the following
experiment was made. The sample holder was kept at
various constant temperatures higher than 2 C, and the
corresponding inner equilibrium pressures were measured. The thermal dependence of the equilibrium
pressure is plotted in Fig. 7 for both NC58 and NC86. It
can be observed that the pressure increases very rapidly
as soon as the temperature rises, suggesting that the
hydrates are very quickly decomposed. This is one
interesting advantage from the storage point of view:
just warming the storage vessel makes the methane
available almost immediately. Moreover, the decomposition of hydrates is so fast that no simple kinetic studies
could be done. Besides, Fig. 7 shows that the slopes of
the curves change abruptly at temperatures close to 7
C. Such temperatures are lower than those for which
hydrates are formed at 8 MPa, i.e., 11 C [8]. However,
this dierence may be explained by the strong hysteresis
always observed between formation and decomposition
of hydrates [6,8,16].

5. Conclusion
The present study leads to the conclusion that
methane storage by hydrate formation in wetted carbonaceous adsorbents is possible, provided that realistic
experimental conditions are fullled. Especially, low
temperatures (close to 0 C) and high pressures (above
34 MPa) are required. Such conditions are serious
disadvantages as far as the application to both natural
gas-powered vehicle and gas transportation is considered. Indeed, the high pressures imply that the storage
vessel must have heavy walls and maybe a cylindrical
shape; thus, the tank could be bulky and heavy, just like
that which is used for compressed gas. Additionally, a
cooling device should be installed. Besides, the weight of
the storage tank should be even enhanced due to the
own weight of the amount of water added to the
adsorbent. Finally, the formation kinetics are so slow
that the simple lling of the tank is a real practical
problem. The only true advantages of such a way of
storing methane seem to be that natural gas is released
very rapidly and almost completely.
The other conclusion of this work is that the classical
carbonaceous adsorbents are not suitable for forming
methane hydrates, because the latter preferentially occur
in very large pores, i.e., macro and mesopores. The
presence of large micropore volumes is thus useless.
Nevertheless, high amounts of methane stored (227 V/V
at 8 MPa and 2 C for Picazine) were found in such
unsuitable adsorbents, and it might be asked if higher
capacities could be reached. For that purpose, purely
macroporous solids should be used. Additionally, their
pore volume should be saturated by water, and the
water should be fully accessible to methane.
This latter assumption was tested using the adsorbent
NC58, which pore space was completely saturated with

200

NC58 (w/c = 1)
NC58 (saturated)

NC58
NC86

Methane uptake (V/V)

10.0

Peq (MPa)

9.5

9.0

8.5

150

100

50

8.0

6 6.2

7 7.4 8

0
9

10

Temperature (C)
Fig. 7. Decomposition of methane hydrates as observed by the increase of pressure during the warming of the storage vessel. Arrows
indicate the dissociation temperatures.

Peq(MPa)
Fig. 8. Comparison between the storage isotherms of the adsorbent
NC58, being either wetted with a weight ratio water/carbon (w/c) 1,
or wetted in such a way that the pore space is saturated by water.

A. Perrin et al. / Carbon 42 (2004) 12491256

water. The corresponding storage isotherm is displayed


in Fig. 8 and compared to that already obtained with the
weight ratio water/carbon 1. It is seen that the new
formation pressure is exactly that of the bulk hydrate
(i.e., 3.18 MPa at 2 C), thus just a little lower than that
quoted in Table 3. Consequently, and as expected, no
huge uptakes can be obtained below the formation
pressure corresponding to bulk hydrates. Besides, the
maximum uptake is almost the same as that which was
found with unsaturated NC58. However, maybe some
adsorbents can be found for which the methane uptake
is so high just above the formation pressure that the
target of 150 V/V is exceeded at 3.5 MPa. Such a result
would be very interesting, because the pressure would be
moderate, allowing various shapes for the storage tank
which walls would not be so thick. Further studies are
thus required. Nevertheless, two serious disadvantages
were observed while the pore space is saturated with
water. First, the equilibria are reached even more slowly
than before; for example, the isotherm of saturated
NC58 given in Fig. 8 required one month and a half.
Secondly, water partially evolves on releasing methane,
and hence the process of llingemptying the storage
tank is not fully reversible.
In the near future, the three other adsorbents will be
saturated with water and studied under the same conditions. Doing this, we aim at correlating the stored
amounts of methane trapped in clathrate hydrates with
meso and macropore volumes.

Acknowledgements
The authors are indebted to Gaz de France for having
brought their attention on methane storage within wet
carbonaceous adsorbents, and for providing several
useful references cited in the present work. Pica France
is thanked for supplying the active carbons.

References
[1] Cook TL, Komodromos C, Quinn DF. In: Burchell, editor.
Carbon Mater Adv Technol. Oxford: Pergamon; 1999. p. 269
302.
[2] Atlanta Gas Light Adsorbent Research Group (AGLARG).
Report to US Dept. of Energy, Contract 466590, 1997.
[3] Kaneko K, Maedo Y, Okui T. Method of storing and transporting gases. Japanese Patent JP 1996 0037526 for Tokyo Gas, 1996.
[4] Tange K. Occluded natural gas storage method. Japanese Patent
JP 2000 161595 for Toyota Motor Corp., 2000.
[5] Zhou L, Li M, Sun Y, Zhou Y. Eect of moisture in microporous
activated carbon on the adsorption of methane. Carbon
2001;39(5):7736.
[6] Zhou L, Sun Y, Zhou Y. Enhancement of the methane storage
on activated carbon by preadsorbed water. AIChE J 2002;
48(10):24126.
[7] Rudzinski W, Everett DH. Adsorption of gases on heterogeneous
surfaces. London: Academic Press; 1992.

1255

[8] Sloan ED. Clathrate hydrates of natural gases. 2nd ed. New York:
Marcel Dekker; 1997.
[9] Smith DH, Wilder JW, Seshadri K. Methane hydrate equilibria
in silica gels with broad pore size distributions. AIChE J 2002;
48(2):393400.
[10] Klauda JB, Sandler SI. Predictions of gas hydrate phase equilibria
and amounts in natural sediment porous media. Marine Pet Geol,
2003;20(5):45970.
[11] Clarke MA, Pooladi-Darvish M, Bishnoi PR. A method to predict
equilibrium conditions of gas hydrate formation in porous media.
Ind Eng Chem Res 1999;38(6):248590.
[12] Uchida T, Ebinuma T, Ishizaki T. Dissociation condition measurements of methane hydrate in conned small pores of porous
glass. J Phys Chem B 1999;103(18):365962.
[13] Kono HO, Narasimhan S, Song F, Smith DH. Synthesis of
methane hydrate in porous sediments and its dissociation by
depressurizing. Powder Technol 2002;122(23):23946.
[14] Miyawaki J, Kanda T, Suzuki T, Okui T, Meda Y, Kaneko K.
Macroscopic evidence of enhanced formation of methane nanohydrates in hydrophobic nanospaces. J Phys Chem B 1998;
102(12):218792.
[15] Seo Y, Lee H, Uchida T. Methane and carbon dioxide hydrate
phase behavior in small porous silica gels: three-phase equilibrium
determination and thermodynamic modeling. Langmuir
2002;18(24):916470.
[16] Perrin A, Celzard A, Mar^eche JF, Furdin G. Unpublished results,
2002.

QUESTIONS AND COMMENTS


A. Linares-Solano, University of Alicante, Spain (linares@ua.es)
Thank you for your very interesting results. You have
clearly shown the limitations of adsorbed methane hydrates as a storage medium for natural gas (low equilibrium, low temperature, heavy tanks etc.) You propose
to improve hydrate formation by using macroporous
carbons rather than a microporous solid. My question is
Why is carbon needed to form methane hydrates,
would not a tank without adsorbent be a better solution
as happens naturally in deep sea water?
A. Celzard
Historically, we started this work due to the very high
storage capacities claimed by the Japanese patents
dealing with methane hydrate formation within porous
carbons, which had to be checked. That is the main
reason why we used carbon to prepare methane hydrates. But the question is very relevant, carbon is
probably not necessary from a chemical point of view;
other porous substrates should also work and maybe
could exhibit higher performances. Anyway, a porous
structure seems to be more favourable for hydrate formation than pure bulk water. Indeed, water occurs in
porous solids in such a way that a high surface of water
is available for gaseous methane. Consequently, preparing hydrates is easier, even if higher formation
pressures are sometimes required due to some additional
capillary pressures. As a comparison, the theoretical

1256

A. Perrin et al. / Carbon 42 (2004) 12491256

limit of 180 V/V which should be reached when 100% of


the water cages are lled with methane is never obtained
at the laboratory scale in pure bulk water; only about
30% of this limit is reached [1 and refs therein], due to
the very low solubility of methane in water. Porous
substrates are thus useful for increasing the eciency of
the hydrate formation process [2], and such a way of
formation is indeed that which naturally occurs in both
permafrost and deep-sea sedimentary rocks.
One advantage or carbon is the fact that physisorption of methane is also possible in the smallest pores.
Thus, the high storage capacities originate both from
hydrate formation and from adsorption, leading to an
optimal lling of the pore space by methane.

References
[1] Link DD, Ladna EP, Elsen HA, Tay CE. Formation
and dissociation studies for optimizing the uptake of
methane by methane hydrate. Fluid Phase Eq.
2003;211(1):110.
[2] Cha SB, Ouar H, Wildeman TR, Sloan ED. A thirdsurface eect on hydrate formation. J Phys Chem.
1988;92(23):64924.
H.P. Boehm, University of Munich, Germany
(Hpb.@uni-muenchen.de)
Taking up Professor Linares question, why should
one use carbon and not macroporous silicas which are
easier to obtain that macroporous carbons?
A. Celzard
Macroporous silicas can be successfully used for
methane hydrate formation, and natural methane hydrates are indeed found in porous siliceous rocks.
However, one major advantage of carbon is its hydrophobic surface (as seen by water adsorption isotherms
[1]), on which water is expected to be much more mobile
than on the surface of an oxide. Indeed, hydrogen
bonding is likely with the latter surface, and hence the
diusion and the reorganization of water molecules required to form clathrates should be more dicult than
on a carbon surface. Consequently, due to the greater
water surface interaction, clathrate formation is expected to be slower in silicas than in porous carbons (all
the other parameters being kept constant).

Reference
[1] Cossarutto L, Zimny T, Kaczmarczyk J, Siemieniewska, T, Bimer J, Weber JV. Transport and
sorption of water vapour in activated carbons. Carbon 2001;39(15),233946.
Valdimir Strelko, Institute for Sorption and Problems
of Endoecology, Kiev, Ukraine (vstrelko@ispe.kiev.ua)
Is it possible that the bad kinetics of formation of
methane hydrates in pores is connected with the diculty of destroying the water structure during the
clathrate formation process?
A. Celzard
The bad kinetics of methane hydrate formation have
several origins. Indeed, water molecules have to rearrange in order to form clathrates. Now, it is well known
that water molecules are strongly associated with each
other, and that the solubility of methane in water is very
low [1], hence clathrate formation is expected to be very
dicult. But the latter is even slower in porous systems
due to percolation eects. Hydrates are rst formed in
the places where water is the most accessible, i.e., in
macropores. While macropores are lled with hydrate
crystallites, higher pressures are required for clathrates
to be formed in smaller pores. Additionally, methane
has to diuse throughout the pores already lled in order to reach the remaining water trapped in smaller
pores. In a connected pore network, the diusion process is thus expected to be more and more hindered as
the gaseous methane has to reach smaller and smaller
pores.
One way of improving the formation kinetics could
be the following. Instead of wetting the carbon and
trying to obtain clathrates by introducing growing
pressures of methane at low temperatures, submitting a
cold dry carbon to a compressed mixture of steam and
methane could lead to clathrates formed with improved
kinetics. At the present time, we have not tested such a
way of proceeding.
Reference
[1] Servio P, Englezos P. Measurement of dissolved
methane in water in equilibrium with its hydrate.
J Chem Eng Data. 2002;47:8790.

Вам также может понравиться