Sie sind auf Seite 1von 7

Colloids and Surfaces A: Physicochem. Eng.

Aspects 357 (2010) 67–73

Contents lists available at ScienceDirect

Colloids and Surfaces A: Physicochemical and


Engineering Aspects
journal homepage: www.elsevier.com/locate/colsurfa

Computational approach to study hydrogen storage in clathrate hydrates


N.I. Papadimitriou ∗ , I.N. Tsimpanogiannis 1 , A.K. Stubos
Environmental Research Laboratory, National Center for Scientific Research “Demokritos”, 15310 Agia Paraskevi, Greece

a r t i c l e i n f o a b s t r a c t

Article history: This work presents collectively, analyzes and compares the results from previous Monte Carlo stud-
Received 30 June 2009 ies performed towards the evaluation of the hydrogen-storage capacity of clathrate hydrates. Clathrate
Received in revised form hydrates can be regarded as a special type of nanoporous material that consists of hydrogen-bonded
16 September 2009
water molecules that form cavities where gas molecules can be entrapped. Three different hydrates
Accepted 1 October 2009
structures (sI, sII, and sH) are examined where hydrogen is either the single guest component or a second
Available online 8 October 2009
substance (promoter) coexists. The storage of hydrogen in hydrates can be simulated as a process of gas
adsorption in a solid material and, consequently, the Grand Canonical Monte Carlo approach seems to
Keywords:
Clathrate hydrate
be the most appropriate approach. These simulations are themselves consistent and, most importantly,
Monte Carlo simulation they are in good agreement with the available experimental data. This fact illustrates the efficiency of
Hydrogen storage the specific computational approach in the study of hydrates. The comparison between the three hydrate
Occupancy types shows that sH hydrates can store larger amounts of hydrogen than sI or sII hydrates. In particular,
the maximum hydrogen content achieved is 3.6 wt% with pure hydrogen, and 1.4 wt% with the use of a
promoter (at 274 K and 500 MPa).
© 2009 Elsevier B.V. All rights reserved.

1. Introduction data. Furthermore, the results of this analysis are used to demon-
strate the GCMC methodology as a powerful tool in the study of
Gas adsorption in solid materials is a research topic with sig- hydrates that can provide significant information complementary
nificant academic and industrial applications [1]. Of significant to the experiment and other theoretical approaches.
interest are the characterization of porous materials [2,3] and gas
storage [4]. While experiments provide valuable results of real 2. Background and methodology
systems, often researchers turn to simulations in order to obtain
complementary information for real systems or examine hypothet- 2.1. Clathrate hydrates
ical ones. Monte Carlo simulations [5,6] have been widely used to
study adsorption of gases in solid porous materials. In this study Clathrate hydrates are crystalline materials that consist of
we are interested in applying the Monte Carlo methodology to the hydrogen-bonded water molecules [7]. Their structural differ-
study of clathrate hydrates. The capacity of hydrates to store large ence from normal ice is that the water molecules of the hydrate
amounts of gases [7] turns them into candidates for the storage of lattice form cavities where molecules of appropriate size can
energy-carrier gases (i.e. methane [8,9] and hydrogen [10,11]). be entrapped. For this reason, clathrate hydrates belong to the
The main objective of this work is to present an integrated eval- class of inclusion compounds. They can be regarded as a spe-
uation of the efficiency of clathrate hydrates in H2 storage by the cial type of nanoporous materials with pores of diameter in the
use of molecular simulations, and in particular, Grand Canonical range 0.7–1.2 nm [7,12]. The stability of hydrates originates from
Monte Carlo (GCMC) simulations. The three most common hydrate the interactions between the enclathrated (guest) molecules and
structures, namely sI, sII, and sH, are examined in a wide pressure the water (host) molecules of the crystal lattice. For this reason,
range (up to 500 MPa) and they are compared in terms of their hydrates are not stable in the absence of the guest molecules while
H2 uptake. In addition, results from different computational stud- they may be stable at conditions where normal ice cannot (e.g.
ies are compared to each other and with available experimental above 0 ◦ C at ambient pressure). There are three common hydrate
structures named sI, sII, and sH. Each structure has different crys-
tallographic properties and contains cavities of different shape and
∗ Corresponding author. Tel.: +30 2106503416; fax: +30 2106525004.
size as shown schematically in Fig. 1. The main properties of these
E-mail addresses: nikpap@ipta.demokritos.gr (N.I. Papadimitriou),
hydrate structures are shown in Table 1.
tsimpano@usc.edu (I.N. Tsimpanogiannis). The first industrial appearance of hydrates was in the produc-
1
Visiting scientist. tion/transportation of natural gas where they were identified as

0927-7757/$ – see front matter © 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.colsurfa.2009.10.003
68 N.I. Papadimitriou et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 357 (2010) 67–73

Table 1
Structure and properties of the three common types of clathrate hydrates. S, M, and L denote the small, medium, and large cavity, respectively.

Hydrate structure sI sII sH


Crystal system Primitive cubic Face-centered cubic Hexagonal
Space group Pm3n Fd3m P6/mmm

Cavity Small Large Small Large Small Medium Large


Description 512 512 62 512 512 64 512 43 56 63 512 68
Cavities/unit cell 2 6 16 8 3 2 1
Cavity radius (Å) 3.95 4.33 3.91 4.73 3.91 4.06 5.71

H2 O/unit cell 46 136 34


Unit cell formula 2S·6L·46H2 O 16S·8L·136H2 O 3S·2M·1L·34H2 O

the cause of pipeline blocking [13]. Since then, however, hydrates first requirement that should be met in order a solid material to be
have been studied for a variety of applications [12,14] including: suitable for practical application in hydrogen storage is high gravi-
energy production from deposits of methane hydrates [15], water metric and volumetric hydrogen content. A wide range of materials
desalination [16], CO2 sequestration [17], and gas separation [18]. has been studied as potentially suitable for H2 storage (e.g. metal
The most common computational tool for the study of hydrates hydrides, carbon-based materials, metal-organic frameworks, etc.)
is a theory proposed by van der Waals and Platteeuw (VDWP) [19] [4].
and subsequently modified by other workers [20–23]. The VDWP Clathrate hydrates have been considered as candidate materials
theory employs principles from statistical mechanics to calculate for hydrogen storage since Mao et al. [10] synthesized hydrogen
the chemical potential of water in the hydrate. The equality of hydrate, which was found to be of the sII structure. This hydrate
chemical potentials of water in the hydrate and in the bulk (liq- structure contains 16 small and 8 large cavities. It was initially esti-
uid or ice) phase allows for phase equilibrium calculations and for mated that each small cavity could contain two H2 molecules and
the determination of the gas content of the hydrate [7]. The basic each large one four. This multiple occupancy of the cavities would
input of the VDWP theory is the potential function within each raise the H2 content of this hydrate up to 5 wt% [10]. However, this
type of cavity that describes the interactions between the guest hydrate is stable only at very high pressures (above 200 MPa at
molecules and the cavity. Moreover, the thermodynamic proper- 280 K [10]), which is far beyond the pressures that could be consid-
ties of the hypothetical empty hydrate are required [24] because it ered acceptable for practical applications (e.g. transportation). The
is used as the reference state. The VDWP theory has been widely high-pressure problem was subsequently resolved with the use of
used to correlate experimental data of the formation conditions of a promoter, which is a substance that occupies the large cavities
hydrates of hydrocarbons during processes in oil and gas industry stabilizing the hydrate at much lower pressures (down to 5 MPa)
with significant success [7]. However, for the most modern appli- [11]. The most common promoter for sII hydrates is tetrahydro-
cations of hydrates including H2 storage, the predictive ability of furan (THF). Obviously, the presence of the promoter reduces the
the VDWP theory is limited. storage capacity of the hydrate since H2 cannot enter the large cav-
ities. Note, however, that some researchers [25,26] have claimed
2.2. Hydrogen storage in clathrate hydrates that if the THF concentration is less than the stoichiometric (i.e. 1
THF molecule per 17 water molecules, which is the stoichiometry
The need to reduce the CO2 emissions and the depletion of oil if all the large cavities are occupied by THF molecules), some large
resources render hydrogen the most suitable fuel, especially for cavities remain empty and H2 molecules can occupy them, there-
the transportation sector. A major issue that needs to be resolved fore, resulting in a material with 4 wt% H2 at moderate pressures
adequately before the establishment of the “hydrogen economy” is [25]. However, this “tuning effect” is not confirmed by recent exper-
hydrogen storage. For economic and safety reasons, hydrogen stor- imental studies [27–30] and there is no consensus in the literature
age in solid materials seems to be the most promising solution. The on this issue.
Binary sH hydrates of H2 have been recently synthesized using
various organic promoters [31,32]. Kim and Lee [33] examined the
possibility of using the binary gas mixture H2 –CO2 to form hydrate
and identified the hydrate formation of structure sI. Linga et al.
[34] conducted experiments for the H2 –CO2 binary hydrate and the
analysis of their results indicated the incorporation of H2 into the
hydrate. In two recent studies Kumar et al. used NMR spectroscopy
[35] and attenuated total reflection IR spectroscopy [36] in order
to examine the issue of cage occupancy for the sI binary H2 –CO2
hydrate. Both studies concluded that CO2 only occupies the large
cavities and the small cavities either are occupied by H2 or remain
nearly empty with a very small amount of CO2 (<2%) in the small
cages. It is evident, therefore, that hydrogen could be enclathrated
in binary hydrates of all the common hydrate structures.

2.3. Simulation details

In this work, we present how Grand Canonical Monte Carlo


Fig. 1. Schematic and notation of the empty water cavities and the resulting hydrate simulations can be used as a computational tool for the study of
structures. The notation 512 stands for a cavity with 12 pentagonal faces. The nota-
clathrate hydrates. A detailed description of the GCMC approach
tion 512 6k (k = 2,4,8) stands for a cavity with 12 pentagonal and k hexagonal faces. The
notation 43 56 63 stands for a cavity with 3 tetragonal, 6 pentagonal and 3 hexagonal
can be found in Refs. [5,6,37]. The most important piece of informa-
faces. tion that can be derived from such simulations is the amount of gas
N.I. Papadimitriou et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 357 (2010) 67–73 69

Table 2
Details of the hydrate structures used in the GCMC simulations considered in this study.

Ref. [38] [39] [40] [41] [42]


Hydrate type sII sII sII sH sI
Water model TIP4P SPC/E SPC/E SPC/E and TIP4P SPC/E
Unit cells 2×2×2 2×2×2 2×2×2 3×3×3 2×2×2
Water molecules 1088 1088 1088 918 368
Lattice constants ˛ = 17.047 Å [10] ˛ = 17.31 Å [44] ˛ = 17.047 Å [10] ˛ = 12.203 Å, c = 9.894 Å [31] ˛ = 12.03 Å [43]
Atomic positions N/A [44] [46] [45,47] [43,48]

that can be stored in a hydrate, as well as the distribution of the gas ties and the guest molecules stabilizing them. This allows for the
molecules among the cavities. GCMC simulations are widely used application of the GCMC methodology. Tanaka [56,57] was the first
in the study of gas adsorption in a solid. The formation of hydrates to apply GCMC in order to calculate the occupancy of several gases
can be represented as an adsorption process where the solid lat- in the cavities of hydrates. In his work, he modified the accep-
tice of the water molecules is the adsorbent and the enclathrated tance criteria in order to be more efficient particularly for hydrates.
gas is the adsorbate. The analysis presented in this work is based Tanaka et al. extended the VDWP theory to account for the multi-
on results from previous works [38–42] from different research ple occupancy of cavities [58]. Sizov and Piotrovskaya [59] used
groups. The GCMC methodology as applied by the aforementioned GCMC to study methane hydrates and extended the methodology
studies cannot be used for phase equilibrium calculations but only to a flexible lattice of water molecules. Methane occupancies with
for the estimation of the gas content of the hydrate. the flexible lattice were lower than those with the rigid one. In gen-
The major input for a GCMC simulation is the geometry of the eral, the comparison showed that the improvement in the results
adsorbent (hydrate crystal lattice) and the force-field that describes with the use of a non-rigid lattice is not worthy the additional
the interactions between the adsorbate (hydrogen) molecules computational cost. This conclusion was confirmed by the work
themselves and between them and the adsorber (lattice of water of Wierzchowski and Monson [60] who combined Monte Carlo
molecules). Table 2 shows how the hydrate structure was con- simulations with several molecular models in order to calculate
structed in each of the works above. The lattice constants and the the free energies and chemical potentials of methane hydrates. In
positions of oxygen atoms have been derived from X-ray diffraction their work, closest agreement with the VDWP theory was achieved
measurements [43–46]. Since the structures are proton disordered, when a rigid lattice was used though the authors attributed this
the positions of hydrogen atoms are selected in such a way that agreement to cancellation of errors. Recently, Katsumasa et al. [38]
minimizes the total dipole moment [47,48]. It seems that there is performed Monte Carlo simulations to calculate hydrogen occu-
a consensus in the selection of the potential function for the H2 pancies in both type of cavities of the sII pure H2 hydrate, while
molecule since the Silvera and Goldman [49] model is exclusively Chun and Lee [39] used GCMC to study the binary H2 -THF hydrate.
used. For the water molecules, two models have been used, SPC/E Papadimitriou et al. studied hydrogen storage in sII [40], sH [41],
[50] and TIP4P [51] but the difference in their interaction param- and sI [42] hydrates and multiple cavity occupancies of sII and sH
eters is very small and for this reason they give almost identical argon hydrates [61] by employing GCMC simulations.
results [41]. The interaction parameters and the partial charges of When calculating cavity occupancies the GCMC simulations
all these force-fields are shown in Table 3. Each run contains about have certain advantages over the traditional VDWP theory. One of
107 random Monte Carlo moves (creation, destruction or displace- the basic requirements of the VDWP theory is that each cavity can-
ment of a molecule). not accommodate more than one gas molecule. On the other hand,
no such assumption is needed for the GCMC simulations and the
occupancy of each cavity is a direct result of the simulation. More-
2.4. Monte Carlo studies of hydrates
over, the VDWP theory involves the interaction between the guest
molecule and the water cavity as a whole. In order to derive a poten-
Tester et al. [52] were the first to incorporate the Monte Carlo
tial function for this interaction, one should smear the interactions
approach in the VDWP theory, in order to accurately calculate
from the water molecules of the cavity over the entire cavity. In the
the potential within the cavities. Other researchers [53–55] uti-
GCMC approach, this approximation is not required and the normal
lized GCMC simulations to evaluate the reliability of Lennard-Jones
molecule–molecule interactions can be taken into consideration.
potential function, for several gas molecules, and calculated the
Langmuir constants.
The formation of hydrates can be considered as a process similar 3. Results and discussion
to the adsorption of a gas in a solid. At thermodynamic equilibrium,
the amount of enclathrated gas at certain conditions depends on 3.1. Pure H2 hydrates
the interactions between host water molecules forming the cavi-
Initially, hydrates where H2 is the single guest component (i.e.
without promoter) are examined. This case sets the upper limit for
Table 3
the H2 content since the addition of a promoter reduces the amount
Parameters of the force-fields used in the GCMC simulations. Point M in the TIP4P
model is a point on the bisector of the H–O–H angle at a distance of 0.15 Å from the of gas that can be enclathrated. For this reason, hydrates of all the
O atom. three common types (sI, sII, and sH) are examined even though
currently there is experimental evidence that H2 by itself can form
Molecule (model used) Atom  (Å) ε (kJ/mol) Charge (e)
only sII hydrate. Note, however, that other small molecules that
H2 O (SPC/E) O 3.166 0.6502 −0.8476 form sII hydrates, such as nitrogen [62] and argon [63,64], can form
H 0.000 0.0000 +0.4238
sH hydrates without the assistance of a promoter, within a specific
H2 O (TIP4P) O 3.154 0.6487 0.0000 pressure range, exhibiting an “sII–to–sH” structural transforma-
H 0.000 0.0000 +0.5200
tion. Therefore, the formation of sI or sH hydrates of pure H2 at
Point M 0.000 0.0000 −1.0400
some P–T range should not be precluded a priori.
H2 (Silvera–Goldman) H 0.000 0.0000 +0.4932 The results from the GCMC simulations [38,40–42] are shown
Center of mass 3.038 0.2852 −0.9864
in Fig. 2. This figure presents the “adsorption isotherms”, i.e. the
70 N.I. Papadimitriou et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 357 (2010) 67–73

Table 4
Large cavity occupancies from GCMC simulations (sII [40], sH [41], and sI [42]). Pressure range and the corresponding dominant number of H2 molecules in the large cavities.

H2 /cavity Pressure range (MPa)

0 1 2 3 4 5 6

Cavity type
512 62 (sI) 0.1–20 20–260 260–500 – – – –
512 64 (sII) 0.1–18 18–125 125–345 345–500 – – –
512 68 (sH) 0.1–10 10–30 30–72 72–147 147–248 248–398 398–500

Fig. 2. Hydrogen content as a function of pressure (“adsorption isotherms”) for pure Fig. 3. Average occupancy of the large cavities for pure H2 hydrates of various
H2 hydrates of various structures obtained from GCMC simulations. structures obtained from GCMC simulations.

average number of guest molecules entrapped in the cavities of this


H2 content as a function of pressure at a constant temperature. type. The “occupancy isotherms” for the large cavities of sI, sII, and
These isotherms have similar forms despite the structural differ- sH hydrates are shown in Fig. 3. This figure presents only the aver-
ences between the three hydrate structures. Below 50 MPa, sI, sII, age occupancy. The probability of finding a specific number of H2
and sH hydrates store about the same amount of H2 . At higher pres- molecules in the large cavity has been discussed in detail in the orig-
sures and up to 500 MPa, sH hydrate seems able to store larger inal works [38,40–42] and will not be discussed any further here. In
amounts of H2 than sI and sII hydrates. This observation confirms Fig. 3, very good agreement is observed between the average occu-
the rough estimate of Strobel et al. [65] that is based on simple cal- pancy of the large cages calculated by Katsumasa et al. [38] and
culations considering the size and number of cavities in each type Papadimitriou et al. [40]. Table 4 shows the approximate pressure
of hydrate. As shown in Fig. 2, for the particular temperature con- ranges for a specific number of H2 molecules to be the dominant
sidered, hydrates cannot store more than 3.6 wt% H2 , which is a average occupancy, at 274 K. As expected, the average occupancy
relatively high value but insufficient for practical applications. Fur- increases with the size of the cavity. This is also demonstrated in
thermore, high pressures are required in order to achieve such high Table 5 where the average occupancy (at 274 K and 500 MPa), as
values of H2 content. In Fig. 2, very good agreement is observed well as the maximum number of hydrogen molecules in the large
between the H2 content calculated by Katsumasa et al. [38] (at cavities is shown. We can observe in Table 5 that the large cavity
273 K), and Papadimitriou et al. [40] (at 274 K). The former work of the sII type can indeed accommodate up to 4 H2 molecules as
uses the TIP4P model for the water molecules while the latter uses has been initially estimated by Mao et al. [10]. However, only 31%
the SPC/E model and there are also some minor differences in the of the large cavities are quadruply occupied at 500 MPa.
GCMC algorithm. The average occupancy results (obtained from GCMC simula-
The overall gas content of a hydrate structure is a function of the tions) for the small cavity are shown in Fig. 4 for sI, sII, and sH pure
number and type of cavities contained in its unit cell. Therefore, in H2 hydrates. Contrary to the initial experimental [10] and compu-
order to better understand the storage capacity of different hydrate tational [66] estimates that reported double occupancy of the small
structures, or possibly explore the use of new hydrate structures, cavity of the sII pure H2 hydrate, the GCMC simulations show that
it is essential to comprehend the average occupancy of all cavity this cavity can accommodate one H2 molecule at most. The fraction
types. The average occupancy of each type of cavity denotes the of doubly occupied small cavities does not exceed 0.2% in any of the

Table 5
Large cavity occupancies from GCMC simulations, at 274 K and 500 MPa (sII [40], sH [41], and sI [42]). Average and maximum numbers of H2 molecules in the large cavities
of the three common hydrate structures.

Cavity type Average number of H2 Maximum number of H2 Fraction of the cavities occupied by
molecules per large cavity molecules per large cavity the maximum possible number of H2
molecules

512 62 (sI) 2.14 3 19%


512 64 (sII) 3.19 4 31%

512 68 (sH) 6.12 7 26.5%


8 2.1%
N.I. Papadimitriou et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 357 (2010) 67–73 71

Fig. 4. Average occupancy of the small cavities for sI, sII, and sH pure H2 hydrates Fig. 5. Hydrogen content as a function of pressure (“adsorption isotherms”) for sI,
obtained from GCMC simulations. sII, and sH binary H2 -promoter hydrates obtained from GCMC simulations.

structures (at pressures up to 500 MPa). As Fig. 4 shows very good promoter-stabilized hydrates and reduces the prospects for their
agreement is observed between the simulations reported by Kat- use in practical applications. In Fig. 5, very good agreement is
sumasa et al. [38] and Papadimitriou et al. [40]. These results are observed between the H2 content calculated by Chun and Lee [39]
in agreement with the most recent experimental measurements and Papadimitriou et al. [40] for the case of H2 -THF binary hydrate.
[29,30,67]. The inability of the small cavity to host more than one The GCMC simulations by Papadimitriou et al. [40] are performed
H2 molecule significantly reduces the hydrogen-storage capacity of at 274 K and the simulations of Chun and Lee [39] at 270 K.
hydrates. For example, in the case of a possible sH hydrate of pure For the cases of the three binary hydrates examined the occu-
H2 , Strobel et al. [31] estimated that the H2 content could reach up pancies of the small cavities are practically the same and are very
to 5.6 wt% (with double occupancy of the small and medium cavi- close to the values for the cases of the pure hydrates. This is clearly
ties and 8 H2 molecules in the large). This value should be compared shown in Fig. 6 where the small cavity occupancy for the pure H2 sII
to the GCMC value of 3.6 wt% [41], which is the result of single occu- hydrate and the binary H2 -THF sII hydrate are shown. This obser-
pancy of the small and medium cavities and 6.2 H2 molecules on vation leads to the conclusion that the presence of the promoter in
average in the large. the large cavities hardly affects the occupancy of the small cavities
As can be seen in Fig. 4, the “occupancy isotherms” of the small although it can have an enormous effect on the hydrate formation
cavity for the three types of hydrates are essentially identical. pressure. Plotted in Fig. 6 are also the GCMC results reported by
This should be an expected observation since the small cavity has Chun and Lee [39] which are in very good agreement with those of
the same shape (512 ) and almost the same size (cavity diameter Papadimitriou et al. [40].
7.82–7.90 Å as shown in Table 1) for all the three hydrate struc- What differs between each hydrate structure is the number of
tures. Therefore, this result indicates that the occupancy in each small cavities per water molecule. There are 0.043 small cavities per
cavity is hardly dependent on the crystal structure beyond the cav- water molecule in sI hydrate while the corresponding values are
ity. The guest molecule mainly interacts with the water molecules 0.118 and 0.147 for sII and sH hydrates, respectively. The medium
of the cavity (first shell of water molecules), while the interactions cavities of sH hydrates are counted as small because their size is
with the more distant shells is negligible. This eliminates the need similar (see Table 1). For binary hydrates of the type H2 -promoter,
for a special treatment of each shell of water molecules, which is a the hydrogen-storage capacity of the three hydrate structures fol-
common practice in some modifications [21] of the VDWP theory. lows the order of the size of the large cavity: sH > sII > sI. We must

3.2. Binary H2 -promoter hydrates

Promoter-stabilized hydrates are cases of great practical inter-


est because a promoter is necessary to either drive towards the
formation of a specific hydrate structure (e.g. for the binary sys-
tem H2 -promoter hydrate we need to select as “promoter” CO2
for sI, THF for sII, or MCH for sH structure to be obtained) or to
reduce the formation pressure of the pure hydrate down to mod-
erate pressures (e.g. THF for sII binary hydrates). In our previous
GCMC simulations we have chosen the following promoters: ethy-
lene oxide (EO) for sI [42], tetrahydrofuran (THF) for sII [40], and
methyl-cyclohexane (MCH) for sH hydrates [41]. The results for
the hydrogen content are illustrated in Fig. 5. As expected, there
is significant difference in the H2 content of the three hydrate
structures. The “adsorption isotherms” have a Langmuir-type form
with a plateau at the value that corresponds to the case where all
the small cavities are occupied by one H2 molecule. These values
are: 0.37 wt% for sI, 1.05 wt% for sII, and 1.40 wt% for sH. Unfor-
Fig. 6. Average occupancy of the small cavities as a function of pressure for the sII
tunately, the inability of the small cavity to accommodate more pure H2 hydrate, and the sII binary H2 -THF hydrate obtained from GCMC simula-
than one H2 molecules limits the hydrogen-storage capacity of tions.
72 N.I. Papadimitriou et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 357 (2010) 67–73

Fig. 7. Hydrogen content for the sII binary H2 -THF hydrate. Comparison between Fig. 8. Radial distribution functions from GCMC simulation of sII pure H2 hydrate
experiments and GCMC simulations. The dashed line (at 1.05 wt%) denotes the case at several pressures and 274 K.
when all small cavities are occupied by a single H2 molecule.

point out that the H2 content when calculated as weight per- shell with diameter of about 2/3 the cavity diameter. The height of
cent depends on the molecular weight of the promoter but this this peak increases with pressure since the population of the dou-
effect is less important than the number of small cavities per water bly occupied cavities also increases. At high pressures triple and
molecule. quadruple occupancy can also occur but neither the position of the
Finally, we evaluate the efficiency of the GCMC simulations in first peak seems to change nor new peaks appear. This means that
the study of the H2 -storage capacity of hydrates by comparing in the case of triple occupancy the H2 molecules form an equilateral
the GCMC results with the available experimental data from sev- triangle with an edge of 2.87 Å. Similarly, in the case of quadruple
eral sources in the literature [27,28,68,69]. The comparison for the occupancy the H2 molecules are placed on the vertices of a nor-
sII binary H2 -THF hydrate is illustrated in Fig. 7. The GCMC sim- mal tetrahedron and the distance between them is 2.87 Å. This is in
ulations by Papadimitriou et al. [40] are performed at 274 K and close agreement with the value of 2.93 Å reported by Lokshin et al.
the simulations of Chun and Lee [39] at 270 K. Some experimental [67], who performed neutron diffraction measurements. The next
measurements in Fig. 7 have been carried out at somewhat lower peak at about 6.2 Å is the superposition of the distances between
temperature because a few degrees of supersaturation (tempera- two neighboring small cavities (6.03 Å), one small and one large
ture lower than the equilibrium temperature at the given pressure) (7.07 Å) and two large cavities (7.38 Å).
always facilitate the experimental procedure. However, this differ-
ence in temperature only slightly affects the hydrogen content and
the comparison can be considered reliable. It is obvious from Fig. 7
4. Conclusions
that there is good agreement between the experimental data and
the GCMC simulations. Furthermore, if one considers the scatter-
We have presented an integrated evaluation of the hydrogen-
ing between the experimental data themselves it can be deduced
storage capacity of clathrate hydrates using Grand Canonical Monte
that experiment and simulation lead to the essentially the same
Carlo simulations. In the pressure range considered (up to 500 MPa)
conclusions regarding the hydrogen content.
and at 274 K, sH hydrates present the higher efficiency in H2 storage
This agreement between experiments and simulations renders
either with or without the use of a promoter. For hydrates without
the GCMC methodology a valuable tool for the study of hydrates
promoter, the maximum H2 uptakes (as wt%) achieved are: 3.5 for
and, especially, in the investigation of the amount of gas that can
sI, 3.3 for sII and 3.6 for sH hydrates. When a promoter is used
be stored in a hydrate. Moreover, recall that the GCMC methodol-
the corresponding values are: 0.4 for sI, 1.1 for sII and 1.4 for sH
ogy has several important advantages relatively to the traditional
hydrates. The small cavities of all the three types are found to be
VDWP theory that make its application very simple and straight-
occupied by one H2 molecule at most while the large cavities of
forward.
the sI, sII, and sH hydrates can host up to 3, 4, and 8 H2 molecules,
respectively. Unfortunately, the promoter-stabilized H2 hydrates,
3.3. Structural analysis although stable at moderate pressures, do not present the desired
hydrogen-storage efficiency to be useful in practical applications.
In this section, we present how the GCMC results can be On the other hand, the hydrates of pure H2 can achieve significant
employed to investigate the configuration of H2 molecules within values of H2 content but very high pressures are required.
the cavities. This analysis is based on the radial distribution func- Apart from the evaluation of hydrates as potential hydrogen-
tion, g(r). This function represents the probability of finding the storage materials, this work demonstrates the GCMC approach as
centers of masses of two H2 molecules at a certain distance r rela- a powerful tool for the study of gas storage in hydrates. The GCMC
tively to the same probability in the ideal gas (with no interactions results from different computational works are in very good agree-
between the molecules). Fig. 8 presents the g(r) derived from GCMC ment with each other as well as with the experimental data from
simulations on an sII hydrate of pure H2 at several pressures [40]. the literature. Contrary to the VDWP theory, the traditional theo-
The first peak at 2.87 Å corresponds to the distance between two H2 retical tool for the study of hydrates, the GCMC approach requires
molecules in the same large cavity. Recall that the diameter of the less assumptions, is more straightforward and can be used in cases
large cavity is 4.73 Å (Table 1). So, the area with the highest prob- where the VDWP theory is not valid (e.g. multiple occupancy of the
ability of finding a H2 molecule in the large cavity is a spherical cavities).
N.I. Papadimitriou et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 357 (2010) 67–73 73

Acknowledgement [31] T.A. Strobel, C.A. Koh, E.D. Sloan, J. Phys. Chem. B 112 (2008) 1885.
[32] A.R.C. Duarte, A. Shariati, L.J. Rovetto, C.J. Peters, J. Phys. Chem. B 112 (2008)
1888.
Partial funding by the European Commission DG Research (con- [33] D.Y. Kim, H. Lee, J. Am. Chem. Soc. 127 (2005) 9996.
tract SES6-2006-518271/NESSHY) is gratefully acknowledged by [34] P. Linga, R. Kumar, P. Englezos, Chem. Eng. Sci. 62 (2007) 4268.
the authors. [35] R. Kumar, I. Moudrakovski, J.A. Ripmeester, P. Englezos, AIChE J. 55 (2009) 1584.
[36] R. Kumar, S. Lang, P. Englezos, J. Ripmeester, J. Phys. Chem. A 113 (2009) 6308.
[37] N. Metropolis, A.W. Rosenbluth, M.N. Rosenbluth, A.H. Teller, E. Teller, J. Chem.
References Phys. 21 (1953) 1087.
[38] K. Katsumasa, K. Koga, H. Tanaka, J. Chem. Phys. 127 (2007) 044509.
[1] S.J. Gregg, K.S.W. Sing, Adsorption, Surface Area and Porosity, 2nd ed., Academic [39] D.H. Chun, T.Y. Lee, Mol. Simul. 34 (2008) 837.
Press, London, 1982. [40] N.I. Papadimitriou, I.N. Tsimpanogiannis, A.T. Papaioannou, A.K. Stubos, J. Phys.
[2] K.S.W. Sing, Adv. Coll. Interf. Sci. 76–77 (1998) 3. Chem. C 112 (2008) 10294.
[3] L.D. Gelb, K.E. Gubbins, R. Radhakrishnan, M. Sliwinska-Bartowiak, Rep. Prog. [41] N.I. Papadimitriou, I.N. Tsimpanogiannis, A.T. Papaioannou, C.J. Peters, A.K. Stu-
Phys. 62 (1999) 1573. bos, J. Phys. Chem. B 112 (2008) 14206.
[4] U. Sahaym, M.G. Norton, J. Mater. Sci. 43 (2008) 5395. [42] N.I. Papadimitriou, I.N. Tsimpanogiannis, A.K. Stubos, Mol. Simul., submitted
[5] M.P. Allen, D.J. Tildesley, Computer Simulations of Liquids, Clarendon Press, for publication.
Oxford, UK, 1987. [43] R.K. McMullan, G.A. Jeffrey, J. Chem. Phys. 42 (1965) 2725.
[6] D. Frenkel, B. Smit, Understanding Molecular Simulation, Academic Press, San [44] T.C.W. Mak, R.K. McMullan, J. Chem. Phys. 42 (1965) 2732.
Diego, 2002. [45] K.A. Udachin, C.I. Ratcliffe, G.D. Enright, J.A. Ripmeester, Supramol. Chem. 8
[7] E.D. Sloan, Clathrate Hydrates of Natural Gases, 2nd ed., Marcel Dekker, New (1997) 173.
York, 1998. [46] M. Yousuf, S.B. Quadri, D.L. Knies, K.S. Grabowski, R.B. Coffin, J.W. Pohlman,
[8] S. Thomas, R.A. Dawe, Energy 28 (2003) 1461. Appl. Phys. A 78 (2004) 925.
[9] A.A. Khokhar, J.S. Gudmundsson, E.D. Sloan, Fluid Phase Equilib. 150–151 (1998) [47] Y. Okano, K. Yasuoka, J. Chem. Phys. 124 (2006) 024510.
383. [48] H. Docherty, A. Galindo, C. Vega, E. Sanz, J. Chem. Phys. 125 (2006) 074510.
[10] W.L. Mao, H.K. Mao, A.F. Goncharov, V.V. Struzhkin, Q. Guo, J. Hu, J. Shu, R.J. [49] I.F. Silvera, V. Goldman, J. Chem. Phys. 69 (1978) 4209.
Hemley, M. Somayazulu, Y. Zhao, Science 297 (2002) 2247. [50] H.J.C. Berendsen, J.R. Grigera, T.P. Straatsma, J. Phys. Chem. 91 (1987) 6269.
[11] L.J. Florusse, C.J. Peters, J. Schoonman, K.C. Hester, C.A. Koh, S.F. Dec, K.N. Marsh, [51] W.L. Jorgensen, J.D. Madura, Mol. Phys. 56 (1985) 1381.
E.D. Sloan, Science 306 (2004) 469. [52] J.W. Tester, R.L. Bivins, C.C. Herick, AIChE J. 18 (1972) 1220.
[12] E.D. Sloan, Nature 426 (2003) 353. [53] V. Natarajan, P.R. Bishnoi, Ind. Eng. Chem. Res. 34 (1995) 1494.
[13] E.G. Hammerschmidt, Ind. Eng. Chem. 26 (1934) 851. [54] K.A. Sparks, J.W. Tester, Z. Cao, B.L. Trout, J. Phys. Chem. B 103 (1999) 6300.
[14] I. Chatti, A. Delahaye, L. Fournaison, J.P. Petitet, Energy Convers. Manage. 46 [55] B. Kvamme, A. Lund, T. Hertzberg, Fluid Phase Equilib. 90 (1993) 15.
(2005) 1333. [56] H. Tanaka, J. Chem. Phys. 101 (1994) 10833.
[15] T.S. Collet, AAPG Bull. 86 (2002) 1971. [57] H. Tanaka, Fluid Phase Equilib. 144 (1998) 361.
[16] H. Kubota, K. Shimizu, Y. Tanaka, T. Makita, J. Chem. Eng. Jpn. 17 (1984). [58] H. Tanaka, T. Nakatsuka, K. Koga, J. Chem. Phys. 121 (2004) 5488.
[17] P.G. Brewer, C. Friederich, E.T. Peltzer, F.M. Orr, Science 284 (1999) 943. [59] V.V. Sizov, E.M. Piotrovskaya, J. Phys. Chem. B 111 (2007) 2886.
[18] S.P. Kang, H. Lee, Environ. Sci. Technol. 34 (2000) 4397. [60] S.J. Wierzchowski, P.A. Monson, J. Phys. Chem. B 111 (2007) 7274.
[19] J.H. van der Waals, J.C. Platteeuw, Adv. Chem. Phys. 2 (1959) 1. [61] N.I. Papadimitriou, I.N. Tsimpanogiannis, A.T. Papaioannou, A.K. Stubos, Mol.
[20] W.R. Parrish, J.M. Prausnitz, Ind. Eng. Chem. Proc. Des. Dev. 11 (1972). Simul. 34 (2008) 1311.
[21] V.T. John, G.D. Holder, J. Phys. Chem. 86 (1982) 455. [62] S. Sasaki, S. Hori, T. Kume, H. Shimizu, J. Chem. Phys. 118 (2003) 7892.
[22] V.T. John, G.D. Holder, J. Phys. Chem. 89 (1985) 3279. [63] H. Hirai, Y. Uchihara, Y. Nishimura, T. Kawamura, Y. Yamamoto, T. Yagi, J. Phys.
[23] J.B. Klauda, S.I. Sandler, Ind. Eng. Chem. Res. 39 (2000) 3377. Chem. B 106 (2002) 11089.
[24] Y.P. Handa, J.S. Tse, J. Phys. Chem. 90 (1986) 5917. [64] A.G. Ogienko, A.V. Kurnosov, A.Y. Manakov, E.G. Larionov, A.I. Ancharov, M.A.
[25] H. Lee, J.W. Lee, D.Y. Kim, J. Park, Y.T. Seo, H. Zeng, I.L. Moudrakovski, C.I. Rat- Sheromov, A.N.A.N. Nesterov, J. Phys. Chem. B 110 (2006) 2840.
cliffe, J.A. Ripmeester, Nature 434 (2005) 743. [65] T.A. Strobel, C.A. Koh, E.D. Sloan, Fluid Phase Equilib. 261 (2007) 382.
[26] D.Y. Kim, Y. Park, H. Lee, Catal. Today 120 (2007) 257. [66] S. Patchkovskii, J.S. Tse, Proc. Natl. Acad. Sci. U.S.A. 100 (2003) 14645.
[27] T.A. Strobel, C.J. Taylor, K.C. Hester, S.F. Dec, C.A. Koh, K.T. Miller, E.D. Sloan, J. [67] K.A. Lokshin, Y. Zhao, D. He, W.L. Mao, H.K. Mao, R.J. Hemley, M.V. Lobanov, M.
Phys. Chem. B 110 (2006) 17121. Greenblatt, Phys. Rev. Lett. 93 (2004) 125503.
[28] R. Anderson, A. Chapoy, B. Tohidi, Langmuir 23 (2007) 3440. [68] K. Ogata, S. Hashimoto, T. Sugahara, M. Moritoki, H. Sato, K. Ohgaki, Chem. Eng.
[29] S. Hashimoto, T. Murayama, T. Sugahara, H. Sato, K. Ohgaki, Chem. Eng. Sci. 61 Sci. 63 (2008) 5714.
(2006) 7884. [69] F.M. Mulder, M. Wagemaker, L. van Eijck, G.K. Kearly, Chem. Phys. Chem. 9
[30] S. Hashimoto, T. Sugahara, H. Sato, K. Ohgaki, J. Chem. Eng. Data 52 (2007) 517. (2008) 1331.

Das könnte Ihnen auch gefallen