Вы находитесь на странице: 1из 10

Chemical Engineering Science 62 (2007) 3930 3939

www.elsevier.com/locate/ces

Characterization of gas hydrates with PXRD, DSC, NMR, and Raman


spectroscopy
Robin Susilo a,b , John A. Ripmeester a, , Peter Englezos b
a Steacie Institute for Molecular Sciences, National Research Council Canada, Ottawa, Ont., Canada
b Department of Chemical and Biological Engineering, University of British Columbia, Vancouver, BC, Canada

Received 15 October 2006; received in revised form 23 March 2007; accepted 30 March 2007
Available online 19 April 2007

Abstract
Structure I (sI) and H (sH) hydrates containing methane were synthesized and characterized with PXRD, DSC, NMR, and Raman spectroscopy.
Three well-known large molecule guest substances (LMGSs) were selected as sH hydrate formers: 2,2-dimetylbutane (NH), methylcyclohexane
(MCH), and tert-butyl methyl ether (TBME). The solid phase analysis conrmed the presence of sH hydrate whenever a LMGS was present.
The presence of a non-hydrate former (n-heptane) did not affect the methane hydrate structure or cage occupancies. Ice to hydrate conversion
was limited when the LMGS amount was less than stoichiometric and synthesized at low methane pressure, but nearly complete conversion was
achieved with temperature ramping and excess LMGS. The methane occupancies were found to depend on the type of LMGS and increased
with pressure. The hydrate with TBME was found to have the smallest methane content followed by the hydrates with NH and MCH. Both
NMR and Raman identied methane and LMGS signals from the hydrate phase, however, the cage occupancy values of sH hydrate can only
be obtained from NMR spectroscopy. The hydrate structures, ice to hydrate conversion, gas content in hydrate and cage occupancy from the
various measurements are consistent with each other.
2007 Published by Elsevier Ltd.
Keywords: Structure H hydrate; Methane; LMGS; Gas storage

1. Introduction
Gas hydrates are non-stoichiometric crystals composed of
water and guest molecules that generally form under high pressure and low temperature conditions. The water molecules serve
as the crystal host framework that is stabilized by the inclusion
of suitably sized guest molecules. There are three well-known
hydrate structures: cubic structure I (sI), cubic structure II (sII)
and hexagonal structure H (sH). Studies on sI and sII hydrate
have been carried out for more than 50 years but not for sH
hydrate, which was just discovered in 1987 at the National Research Council of Canada (NRC) (Ripmeester et al., 1987). In
recent years, the study of sH hydrates has emerged rapidly because of their unique properties. Firstly, two guest molecules
of different sizes are required for sH hydrate, unlike sI and sII
where one guest molecule is enough to stabilize the crystal.
Corresponding author. Tel.: +1 613 993 2011; fax: +1 613 998 7833.

E-mail address: John.Ripmeester@nrc-cnrc.gc.ca (J.A. Ripmeester).


0009-2509/$ - see front matter 2007 Published by Elsevier Ltd.
doi:10.1016/j.ces.2007.03.045

Secondly, sH hydrate has the largest cage among all the known
hydrate structures hence larger molecules like methylcyclohexane can t into the cavity. The addition of a large molecule
guest substance (LMGS) may also reduce the equilibrium pressure while maintaining high gas storage capacity due to the accessibility of both small and medium cages. Hence sH hydrate
is seen as a valid potential and attractive opportunity for gas
storage application (Englezos and Lee, 2005; Khokhar et al.,
1998; Tsuji et al., 2005; Mori, 2003). However, a solid foundation on which to base sH hydrate characterization is vital
before technological developments can be pursued.
The study on a molecular scale is essential to make sure
that the hydrate properties are known prior to advancing to
a larger synthetic scale. Powder X-Ray diffraction (PXRD),
Raman and Nuclear Magnetic Resonance (NMR) spectroscopy
are the well-known tools used for solid structural analysis,
including that of gas hydrate (Ripmeester and Ratcliffe, 1988,
1999; Sloan, 2003; Tulk et al., 2000). Such techniques are
required to obtain information on crystal structure, crystal

R. Susilo et al. / Chemical Engineering Science 62 (2007) 3930 3939

dimension/volume, and composition/cage occupancy values.


A thermal analysis technique such as Differential Scanning
Calorimetric (DSC) is also employed (Dalmazzone et al., 2002;
Giavarini et al., 2006; Parlour et al., 2004). However, a combination of all instrumental methods to characterize the hydrate
phase has not been reported so that the information obtained
from each technique has not been used without relating to the
others to obtain a more complete model. Interpreting the experimental data from PXRD and NMR for methane hydrate in
all hydrate structures is already well established and often the
measurement from each instrumental method may explain the
results independently. However, Raman spectroscopy still requires attention especially in determining hydrate structure and
the composition for complex systems like sH hydrate.
Both Raman and NMR spectroscopy have been successfully
employed to study gas hydrate at a molecular level (Kini et
al., 2004; Komai et al., 2004; Pietrass et al., 1995; Yoon et
al., 2004). These instrumental methods are based on different
principles and hence have different advantages and limitations.
Raman is convenient for in situ measurements but is not so
diagnostic for structure and composition whereas NMR is more
sensitive to changes in local environment and is quantitative
in terms of numbers and speciation, hence of greater value.
A direct comparison between the two is required to establish
a common understanding of the interpretation of the spectra.
Both NMR and Raman gave comparable cage occupancy values for methane in structure I (sI) hydrate (Uchida et al., 2003;
Wilson et al., 2002) yet a discrepancy was reported for methane
in mixed hydrate with carbon dioxide (Wilson et al., 2002).
Methane occupancy in sH hydrate also has been reported from
NMR and single crystal diffraction study (Seo and Lee, 2003;
Udachin et al., 2002a,b). However, it is still unsure if
Raman can also be used to obtain such occupancy values for sH
although the identication of methane signals has been reported
(Chou et al., 2000; Sum et al., 1997; Sun et al., 2005; Uchida
et al., 2006) and occupancy values were suggested (Uchida
et al., 2006). Due to the different chemical environment, both
Raman and NMR show that when a molecule is encaged the signal in the hydrate phase may be shifted from those in the liquid
or gas phases. For the encaged methane molecule, its signature
in all three hydrate structures is shifted to a lower frequency
in Raman (Sum et al., 1997) and lower eld in NMR (Seo and
Lee, 2003; Subramanian et al., 2000). Generally the shifts in
Raman spectra are not structure specic and a large cage/small
cage occupancy ratio has to be derived from other considerations. Hence one has to be more careful in determining hydrate
structure and its composition from Raman. The 13 C methane
chemical shift from NMR spectroscopy is more structure specic so that there is rather less ambiguity. However, the chemical shift of methane in the small cages of all hydrate structures
is quite similar due to the similar chemical environment and
the signal of methane in 512 and 43 56 63 cages of sH hydrate
can be expected to be very near to each other. Surprisingly, it
was reported recently that Raman signals for methane could be
assigned to both contributions from both small cages (Uchida
et al., 2006). Most Raman spectra reported previously for sH
hydrates in the literature have shown only one broad peak at

3931

around 29132918 cm1 (temperature dependent) that corresponds to methane in both small and medium cages (Chou et
al., 2000; Sum et al., 1997; Sun et al., 2005). None of this work
shows a distinct peak for the two except the last published
spectra. So it is interesting to investigate this puzzle.
The objective of this study is to present a direct comparison of the three solid-state analytical tools (PXRD, NMR,
and Raman) plus DSC for sH hydrate characterization. In addition, the presence of n-heptane (nC7 ) as a non-hydrate former will also be studied along with a system without LMGS
added. The structural information from PXRD will be linked
to both NMR and Raman shifts. The signals from NMR and
Raman will be used to identify the corresponding molecules
in each phase and to determine the cage occupancy. The occupancy values obtained from the spectroscopic techniques will be
compared with the gas content measured by decomposing the
hydrate.
2. Experimental apparati and procedures
Hydrates were synthesized from freshly ground ice particles
that were poured by gravity into a 50 ml pressure vessel. Approximately 10 g of ice powder was used, with the LMGS
sprayed on top of the ice with a syringe after loading the ice
powder. The amount of LMGS was varied from 200% (excess)
to 50% of the stoichiometric composition. The list of chemicals/LMGS used in this study is summarized in Table 1. The
loading procedure was performed in a freezer at 253 K to
prevent the melting of the ice. The vessel was then immersed
in a constant temperature bath containing a watermethanol
mixture and connected to a valve and pressure transducer. The
zero-time of the measurements was recorded as the gas pressure vessel reached the desired value. All measurements were
performed at 253 K over about 20 h. At the end of the 20-h
period the temperature was increased to a point above the icepoint (274 K) within 5 min in order to enhance the conversion
of ice into hydrate. It is well known that temperature ramping enhances the conversion to hydrate (Wang et al., 2002).
Two starting pressure conditions were chosen, 43 (Low) and
81 bars (High) which would give nal pressures well below and
above the equilibrium value for sI methane hydrate at 274 K.
The hydrate samples were collected at liquid nitrogen temperature (82 K) at the end of the experiment after signicant pressure drops were no longer observed (almost complete hydrate
conversion achieved). The recovered hydrate samples were kept
in liquid nitrogen for subsequent analysis.
Table 1
List of chemicals used in this work
Chemical

Certied purity

Supplier

Methane
tert-Butyl methyl ether (TBME)
Neohexane (NH)
Methylcyclohexane (MCH)
n-Heptane (nC7 )
Water

UHP grade
99.9%
99%+
99%+
99%+
Distilled

Praxair
Sigma Aldrich
Sigma Aldrich
Sigma Aldrich
Omnisolv

R. Susilo et al. / Chemical Engineering Science 62 (2007) 3930 3939

Crystal structures and lattice constants were obtained from


PXRD. The PXRD patterns were recorded at 85 K on a
Rigaku Geigerex diffractometer ( = 1.79021) in the /2
scan mode. The XRD experiments were carried out in step
mode with a xed time of 5 s and a step size of 0.05 for
2 = 550 with a total acquisition time of 75 min for each
hydrate sample. The crystal structure was then correlated to
results from 13 C magic angle spinning (MAS) NMR at 193 K
and Raman spectroscopy at liquid nitrogen temperature. A
Bruker DSX 400 MHz NMR spectrometer was used to analyze
the hydrate structure and cage occupancies of methane and
LMGS. Prior to the NMR measurements, hydrate samples were
ground under liquid nitrogen and packed in a 7 mm zirconia
rotor, which was loaded into a variable temperature probe. All
13 C NMR spectra were recorded at 2 kHz spinning rate. A
single pulse excitation (90 of 5 s) and pulse repetition delay
of 300 s under proton decoupling were employed. The crosspolarization technique was also employed to distinguish the
signals arising from the solid phase from those of the liquid
phase. Adamantane was used as the external chemical shift reference at 298 K and 38.56 ppm. An Acton Raman spectrograph
with ber optics and equipped with a 1200 grooves/mm grating
and a CCD detector was used in this study. An Ar-ion laser was
used as the excitation source emitting at 514.53 nm. The laser
was focused on the sample by a 5 microscope objective. The
spectrograph was controlled with a computer and the spectra
were recorded with a 1 s integration time over 530 scans. All
spectra were referenced to methane gas at 2918 cm1 .
The amount of hydrate, unreacted ice and unreacted LMGS
was determined by DSC. Hydrate samples were sealed in an
aluminum pan under liquid nitrogen and placed in the DSC cell
at 150 C. The sample was equilibrated at that temperature for
10 min before heating to 30 C with the rate of 5 C/ min. The
samples were all frozen at 150 C so that a phase change due
to melting appeared at the melting point of the LMGS, hydrate
and ice. The areas for each phase transition peak correspond
to the energy absorbed due to decomposition or melting. The
amount of the corresponding phase can be calculated by calibrating the peak area with the heat of fusion values given in the
literature (Anderson, 2004; Linstrom and Mallard, 2005). The
dissociation enthalpy of hydrate was calculated from the hydrate phase equilibria data using the ClausiusClapeyron equation (Yoon et al., 2003). Hence the amount of unreacted ice and
LMGS can be calculated from the mass balance. The hydrate
conversion and the gas stored in the hydrate were also crosschecked by measuring the gas release while decomposing the
hydrate in a vacuum line of a known volume.
3. Results and discussions
All synthetic hydrate samples were rst analyzed with PXRD
to conrm the presence of hydrate and to verify the crystal
structure. A typical PXRD pattern of sH hydrate is shown in
Fig. 1. The pattern was tted to a standard sH hydrate pattern
(space group P6/mmm) to obtain the lattice constants and unit
cell volumes, which are summarized in Table 2. The unit cell
volumes were found to increase slightly on going from TBME,

3
0
1

2
0
1

Intensity

3932

1
0
2
1
0
1
10

0
0
2
20

2
1
2

2
2
0
2
1
0
*

2 0
12 0
10 3
2
*

3
1
0

2
0
3

1
0
3
*

30
2 [Degrees]

3 2
2 2 1
0 2 4
4

22
21 3 3
23 1 0
23
40

50

Fig. 1. PXRD pattern of sH hydrate (TBME + methane).

Table 2
Lattice constants and unit cell volumes of synthesized hydrate obtained at
82 K
System

Hydrate
structure

Lattice

constants (A)

Unit cell
3)
volume (A

CH4 + H2 O
CH4 + H2 O + nC7
CH4 + H2 O + TBME

I
I
H

1665
1665
1294

CH4 + H2 O + NH

CH4 + H2 O + MCH

a = 11.85
a = 11.85
a = 12.16
c = 10.10
a = 12.18
c = 10.08
a = 12.16
c = 10.13

1295
1297

NH, to MCH. The presence of ice is indicated by the asterisk


in Fig. 1. The intensity of the ice peaks for the sample corresponded to the amount of unreacted ice and some ice resulting
from condensation of moisture in the air, which was found to be
minimal for most samples. For those hydrates that were synthesized with an amount of LMGS that was less than stoichiometric, the conversion of hydrate was very low at low pressure (sH
stability region) as the intensity of ice peaks was predominant.
Some ice particles that were not exposed to LMGS formed sI
hydrate at higher pressures. Consequently, sI and sH hydrate
were detected by PXRD. Hydrates synthesized without LMGS
or n-heptane as a non-hydrate former form sI hydrate. Once the
PXRD conrmed that the solid phase contained hydrate, the
gas stored in the hydrate phase was measured and the sample
was further analyzed to obtain the hydrate composition and unreacted ice content by NMR, Raman spectroscopy, and DSC.
Table 3 summarizes the amount of gas stored in the hydrate
as measured by decomposing the hydrate under vacuum. The
volume ratio (v/v) of gas stored in the hydrate phase that is
relevant to practical interests is reported in the table. The ratio
of gas per volume of water generated upon decomposition is
also given, which is slightly higher due to the higher density
of water than the hydrate phase. Synthetic hydrate made with
n-heptane or without LMGS store the same amount of methane

R. Susilo et al. / Chemical Engineering Science 62 (2007) 3930 3939

3933

Table 3
Amount of gas stored in hydrate together with nal hydrate conversion achieved
LMGS amount/system

Pressure

[v/v] Gas/hydrate

[v/v] Gas/water

Conversiona (%)

No LMGS
200% nC7
50% TBME

High
High
Low
High
Low
High
Low
High
Low
High
Low
High
Low
High

173
173
32
131c
103
125
20
144c
130
139
72c
157c
132
142

210
210
41
166
131
160
26
184
166
177
92
200
168
181

85
85
22b
98d
99
96c
14b
96d
98
91
45b,d
99d
98
87

200% TBME
50% NH
200% NH
50% MCH
200% MCH
a The

conversion was calculated from the total methane uptake divided by the uptake if all ice/water is converted into hydrate taking methane occupancies
into account.
b The conversion was obtained from the gas content measurements.
c The hydrate phase was calculated using lattice constants from sH hydrate.
d The conversion was calculated from total methane uptake divided by the uptake if all ice/water is converted into hydrate taking methane occupancy into
account for both sI and sH hydrate.

Temperature [C]
-140 -120 -100 -80
0
-10
Heat Flow [mW]

gas (173 v-gas/v-hydrate). The addition of LMGS lowers the


hydrate formation pressure by forming sH hydrate so it is favorable for methane storage and transport application, however, as
a result the gas storage capacity in the hydrate phase decreases.
The amount of methane gas stored in sH hydrates was found to
be approximately 2040% lower (103142 v-gas/v-hydrate, depending on the methane occupancy) than in sI hydrate. This is
due to the accessibility of all (both small and large cages) of sI
hydrate towards methane whereas only the small and medium
cages are accessible for sH hydrate. A methane molecule does
not ll the large cavity of sH hydrate effectively, so the LMGS
is preferred as cage occupant. The gas stored in sH hydrate depends on the type of LMGS used and on the pressure at which
the hydrate is synthesized. The systems with MCH and NH
have higher methane content than TBME. A higher hydrate formation pressure increases the methane storage for all systems.
Hydrate synthesized with 50% LMGS at lower pressures has
the least methane content due to limited hydrate conversion.
At higher pressures, the methane content increases due to the
formation of sI hydrate.
Fig. 2 shows a DSC plot obtained from melting sH hydrate in
the presence of MCH. Similar plots were obtained for the other
hydrate forming systems. The melting curve obtained shows
that three distinct phases melt at different temperatures. All
phases are solid at the initial temperature of 150 C. As the
temperature was increased, the unreacted LMGS melted rst
followed by hydrate and nally ice. The amount of unreacted
LMGS, hydrate, and ice (in moles) was calculated by calibrating the peak areas with the corresponding heat of fusion. The
relevant heat of fusion data and hydration numbers are given
in Table 4. The rst endothermic peak appeared at the melting
points of LMGS, which were at 126, 109, and 100 C for
MCH, TBME, and NH, respectively. Hydrate started to dissociate at the corresponding hydrate equilibrium temperatures

Excess MCH
(282 mJ)

-60

-40

-20

20

40

sH hydrate
(3200 mJ)

-20
-30
-40

Ice (6831 mJ)


Unreacted ice
plus dissociated
from hydrate

-50
-60
Fig. 2. DSC melting curve of (methane + MCH) sH hydrate.

at 1 atm, which were approximately 70, 60 and 55 C


for TBME, MCH, and NH, respectively. The ice melting peak
that appeared at 0 C resulted from unreacted ice and ice from
hydrate decomposition. Therefore, the ice that originated from
the dissociated hydrate needed to be determined rst to obtain the amount of unreacted ice in the sample and hence the
hydrate conversion. This was obtained by multiplying the
amount of dissociated hydrate with the hydration number which
was acquired from NMR measurements. It depends on the cage
occupancy and varies between 5.67 and 7.00. The amount
of unreacted ice was then calculated by subtracting the total
amount of melted ice from that of the hydrate. The peak area
and the corresponding amount of each phase from Fig. 2 are
given in Table 5 for the MCH system synthesized at lower
pressure. The hydrate conversion for that particular sample
was found to be 97%. This number is comparable to the total
conversion value calculated from the gas uptake taking into

3934

R. Susilo et al. / Chemical Engineering Science 62 (2007) 3930 3939

Table 4
Heat of fusion and hydration number of LMGS, ice, and sH hydrate
Heat of fusion
(kJ/mol)

Hydration numbera

TBME
NH
MCH
Ice
(CH4 + TBME) hydrate
(CH4 + NH) hydrate
(CH4 + MCH) hydrate

7.60
0.58
6.75
6.01
16.21
18.47
17.93

7.00b
6.30b
6.18b

Intensity

System

High power decoupling


Cross-polarization

6.19c
5.71c
5.67c

a Hydration number was calculated from the cage occupancy values obtained
by NMR.
bValues reported correspond to the hydrate system synthesized at low
pressure with LMGS excess (200%).
cValues reported correspond to the hydrate system synthesized at high
pressure and LMGS excess (200%).

50

40

0.032

44

1.137

1.105

97

-20

Intensity

0.042
0.179

-10

High-power decoupling
Liquid TBME

Integrated Amount from Amount from Conversion into


area (mJ) melting
hydrate
hydrate (%)
(mmol)
(mmol)

MCH
282
(CH4 +MCH) 3200
hydrate
Ice/water
6831

20
10
0
Chemical Shift [ppm]

Fig. 4. 13 C MAS-NMR spectra of (methane + MCH) sH hydrate, synthesized


with 200% MCH at low pressure condition (Final pressure 17 bars at 274 K).

Table 5
DSC summary of methane + MCH hydrate formed at low pressure with
excess LMGS
Phase

30

High powerdecoupling
Cross-polarization

80

60

40

20

-20

Chemical Shift [ppm]

Intensity

Fig. 5. 13 C MAS-NMR spectra of (methane+TBME) sH hydrate, synthesized


with 200% TBME at low pressure condition (nal pressure 21 bars at 274 K).

No LMGS (sI)
NH (sI+sH)
50

40

30

20

10

-10

-20

Chemical Shift [ppm]

TBME (sI+sH)
TBME (sH)

Fig. 3. 13 C MAS-NMR spectra of (methane + NH) sH hydrate, synthesized


with 200% NH at low pressure condition (nal pressure 17 bars at 274 K).

account the cage occupancy and also from directly measuring


the gas stored in the hydrate phase, as given in Table 2.
Figs. 35 show 13 C MAS NMR spectra obtained at 193 K
with 2 kHz spinning rate. The spectra were acquired by crosspolarization/high power proton decoupling program to distinguish the solid phase signal from that of the liquid (excess
LMGS) and hydrate phases, so the signals acquired with crosspolarization refer to the hydrate (solid) phase only. Fig. 6 shows
the magnied spectra in the methane region for sI, sI + sH,
and sH. The signature of the methane signals from sH hydrate

MCH (sH)
NH (sH)
0

-2

-4

-6

-8

-10

-12

Chemical Shift [ppm]


Fig. 6. 13 C MAS-NMR spectra around methane region obtained at 193 K.

is obvious at around 4.4 ppm. The sample with 50% LMGS


formed at high pressure showed the appearance of another
methane signal at 6.5 ppm. This is the signature of methane

R. Susilo et al. / Chemical Engineering Science 62 (2007) 3930 3939


Table 6
Chemical shift of methane and LMGS in the hydrate and liquid phase obtained from
System

CH4 hydrate
(CH4 + TBME) hydrate

13 C

3935

NMR calibrated by adamantane at 298 K

Hydrate phase

Liquid phase

Small cage

Medium cage

Large cage

4.10
4.30

4.70

6.50
26.91
48.41
73.14
8.53
29.20
30.27
36.91b
23.49
27.24b,c
33.71b

(CH4 + NH) hydrate

4.40a

(CH4 + MCH) hydrate

4.40a

36.08b

27.29
49.37
72.47
9.60
29.46
31.04
36.95
23.88
26.95
27.31
33.72
36.09

a Methane

signals from the small and medium cages overlap.


chemical shift for these signals overlapped with the liquid phase (fairly close). The number given is just approximated. The intensity of these signals
was assigned by the corresponding molecular formula with respect to other distinguishable signal, generally from the methyl group (8.53 ppm for NH and
23.49 ppm for MCH).
c MCH signals from the hydrate phase at around 27 ppm are too close to be separated.
b The

in the large cage of sI hydrate, as seen from pure sI methane


hydrate (no LMGS) signals. The methane signal from the small
cage at 4.1 ppm is not clearly seen because it overlaps with
sH signals.
Methane molecules in the small and medium cages of NH and
MCH systems were not distinguishable because of peak overlap. Hence the occupancies of methane in the small and medium
cages for NH and MCH system are reported as the numberweighted average occupancy for the two cages. Interestingly,
the previously reported NMR spectra did show doublets even
for the NH (Seo and Lee, 2003) and MCH (Yeon et al.,
2006) systems whereas a resolved doublet was not seen in this
study. This can be ascribed to the difference in temperature
(50 K) used for the measurements, reported here and those of
Seo and Lee (2003) and Yeon et al. (2006). We used a low
temperature of 193 K in this study to prevent hydrate decomposition during data acquisition in the non-pressurized sample
tubes; the recording of data took a minimum of an hour to obtain a good signal to noise ratio because of long 13 C relaxation
times and the low natural abundance of 13 C in the hydrate
samples.
The dynamics of the guest and host molecules in the hydrate lattice play a key role in determining the maximum resolution obtainable at any temperature (Collins et al., 1990). The
sharpest lines are obtained when the water motions are fast as
then the hydrate cages have their proper high crystallographic
symmetry and the chemical shifts are sharp. If the water motions are slow, the guests do not access all parts of the cage and
the chemical shifts are somewhat distributed and the resonance
lines are broader. For the three guests studied here, the water
dynamics are fastest for TBME (a guest containing oxygen can
inject Bjerrum defects into the lattice) and hence the resonance
lines are sharp at low temperature even for sH hydrate containing TBME (4.3 and 4.7 ppm), but not for hydrates containing
NH or MCH (Collins et al., 1990; Ripmeester and Ratcliffe,

1999). Other evidence for the guest-dependent dynamics comes


from dielectric measurements (Davidson, 1972) and reaction
rates for ice with polar and non-polar guests (Gulluru and
Devlin, 2006), and it is a key feature in the NMR spectroscopy
of clathrate hydrates (Ripmeester and Ratcliffe, 1996). This explains the differences observed for the 13 C NMR spectra of
methane in sH hydrate in this study and that of Seo and Lee
(2003) and Yeon et al. (2006). The intensity ratio of the signals
at 4.3 and 4.7 ppm is approximately 3:2, which corresponds
to methane in the small and medium cages.
The summary of chemical shifts from methane and LMGS in
the hydrate and liquid phases are summarized in Table 6. The
LMGS signals from the hydrate phase are generally shifted to
higher eld from the corresponding liquid signals. A few exceptions were observed where the signal did not shift much from
the liquid signals (MCH), or, where it is shifted to lower eld
in the case of the TBME resonance at 73 ppm. The intensity
of the LMGS signals from the hydrate phase that were indistinguishable from that of the liquid phase was assigned from
the corresponding molecular formula and the other signals that
were distinguishable, generally those from the methyl groups.
The intensity ratio of methane and LMGS peaks from the hydrate phase was then used to calculate the cage occupancy after
taking into account the number of molecules per cage and carbon atoms per molecule. It is generally safe to assume that the
large cage of the hydrate has to be completely lled to maintain hydrate stability. Hence the determination of the methane
occupancy was taken directly from the intensity ratio. For the
TBME system, the cage occupancy was calculated from the
statistical thermodynamics equation for the hydrate phase:
ow =

RT
[3 ln(1 S ) + 2 ln(1 M ) + ln(1 L )].
34

The value of ow was 1187.5 J/mol (Mehta and Sloan, 1996).

R. Susilo et al. / Chemical Engineering Science 62 (2007) 3930 3939

Table 7
Cage occupancy values obtained by

13 C

MAS NMR

LMGS amount/system

Pressure

Hydrate
structure

S

No LMGS
200% nC7
50% TBME

High
High
Low
High
Low
High
Low
High
Low
High
Low
High
Low
High

I
I
H
Ha
H
H
H
Ha
H
H
Ha
Ha
H
H

0.87

0.87

0.78
0.74
0.92
0.63
0.78
0.76
0.97
0.79
0.84b
0.94b
0.88b
0.99b
0.91b
0.99b
0.90b
1.00b

200% TBME
50% NH
200% NH
50% MCH
200% MCH

M

L
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00

a The presence of methane in sI hydrate was also seen. The intensities of the
methane in the small cage of sI and sH hydrate were overlapped, hence the
contribution from sI was calculated with the assumption that its occupancy
was the same as sI methane hydrate by itself and subtracted from the total
amount. The occupancy values which belong to sH hydrate only are reported.
b The spectra from NMR do not distinguish the methane in the small and
medium cages due to very close chemical shift and broader peak than TBME
system. Hence the average occupancy from the two cages is reported.

All cage occupancy values are summarized in Table 7. The


methane content of the hydrates was found to increase with
pressure for all systems and the methane cage occupancy was
found to vary with the LMGS and hydrate formation pressure
in the following order: MCH (0.90), NH (0.88), and TBME
(0.77). This is interesting because the rate of hydrate formation obtained from the kinetics experiments through gas uptake
and spectroscopy indicated the opposite trend (Lee et al., 2005;
Susilo et al., 2006). Thus, reaction rates and methane content
of hydrates are not correlated with each other, as one indeed
would expect the rst being related to kinetics, the second to
thermodynamics. However, the reaction rates were correlated
with the solubility of the LMGS in the water phase (Susilo
et al., 2005). The lower methane occupancy for the TBME system may be associated with a lower net driving force present
during the hydrate synthesis. The driving force here is dened as the difference between the experimental and equilibrium pressure at a given temperature. One factor may be that
TBME has more water solubility than NH and MCH, and since
water soluble species will change the activity of the water
they will show some activity for hydrate inhibition. Since all
hydrates were synthesized under the same pressure and temperature conditions, the most stable system with the lowest equilibrium pressure has the highest driving force. Among all system
investigated, NH and MCH systems are much more stable than
TBME system. Hence the methane occupancies for the NH and
MCH hydrates are higher than for the TBME system. Although
it is also important to note that the occupancy values reported
here are based on non-isobaric and non-isothermal hydrate synthesis this is not seen as playing a major role in the results.
The methane occupancy for the MCH system was also reported from a single crystal study (Udachin et al., 2002a,b). The

reported occupancy was 0.80, which is lower than the values


measured in this study, but this is due to the lower driving force
used during the synthesis of the crystals for diffraction. The
methane occupancies obtained from NMR are also consistent
with the amount of gas released by decomposing the hydrates.
However, our measured gas content is signicantly lower than
the values suggested previously (Khokhar et al., 1998) and in
fact it is difcult to understand how the latter values arose as
there seems to be no fundamental basis for these values based
on structure and maximum cage occupancies. However, it is
clear that one has to consider all factors in selecting the best
LMGS for gas storage applications, namely the hydrate stability conditions, the gas storage capacity, and the kinetics of
formation and decomposition.
In the stability region of sI hydrate, the stable structure is sH
hydrate when excess LMGS is present. A mixture of sI and sH
hydrate was encountered when less LMGS was used, as seen
in Fig. 6, although the amount of LMGS did not seem to affect
the methane occupancy. At this point it is clear that 13 C MAS
NMR is quite capable of identifying hydrate structure and cage
occupancy in sH hydrate and systems that may contain sI/sH
hydrate mixtures. The hydration number can be calculated from
the occupancy values, as given in Table 4. The chemical shift
of methane is specic and clear for each hydrate structure and
the intensity ratio can be used to determine the cage occupancy.
However, the signals for methane in the small cages for all
hydrate structures are in the same vicinity and hence it can
be difcult to separate them accurately when more than one
structure is present.
Hydrate samples were also analyzed with Raman spectroscopy, as shown in Figs. 710. There are many signals
observed in the Raman spectra, however, those most intense
are those in the CH and OH bond stretching regions between
2800 and 3500 cm1 . Unfortunately, the CH stretching signals from methane and the LMGS often will overlap although
the intensity of the methane peak can be expected to be much
higher than those arising from the LMGS. Fig. 11 shows the
magnied spectra in the methane region. There is only a single,
strong CH stretch signal at 2913 cm1 that corresponds to
methane in sH hydrate. No peak splitting was seen as could be
attributed to methane in the two cages of sH hydrate. On the

MCH liquid
Methane +MCH (sHhydrate)
Intensity

3936

2800

2900

3000

3100

3200

3300

3400

3500

Wavenumber [cm-1]
Fig. 7. Raman spectra of (methane + MCH) sH hydrate at 85 K.

R. Susilo et al. / Chemical Engineering Science 62 (2007) 3930 3939

3937

NH liquid
Methane+NH (sH hydrate)
Intensity

No LMGS(sI)
TBME (sI+sH)
TBME (sH)
MCH (sH)
NH (sH)

2800

2900

3000

3100

3200

3300

3400

3500

Wavenumber [cm-1]

2880

Fig. 8. Raman spectra of (methane + NH) sH hydrate at 85 K.

2890

2900

2910

2920

2930

2940

2950

Wavenumber [cm-1]
Fig. 11. Raman spectra around methane region obtained at 85 K.

Intensity

TBME liquid
Methane+TBME hydrate (sHhydrate)

2800

2900

3000

3100

3200

3300

3400

3500

Wavenumber [cm-1]
Fig. 9. Raman spectra of (methane + TBME) sH hydrate at 85 K.

Intensity

TBME liquid
Methane+TBME (sI+sH) hydrate

2800

2900

3000

3100

3200

3300

3400

3500

Wavenumber [cm-1]
Fig. 10. Raman spectra of (methane + TBME) sI + sH hydrate at 85 K.

other hand, the 13 C methane NMR signal in TBME hydrate


does show such splitting. Another methane signal appears at
2904 cm1 for the hydrate formed at high pressure and low
(50%) LMGS. The intensity ratio for the signals at 2913 and
2904 cm1 is 3:2, which coincides with the ratio of small to
medium cages. One could assign this second line at 2904 cm1
as also arising from methane in the medium cage of sH hydrate.
However, the PXRD and NMR measurements do not support
this, suggesting that the sample is a mixture of sI and sH hydrate
with the signal at 2904 cm1 arising from methane in the large

cage of sI hydrate. Uchida et al. (2006) argued that indeed two


methane signals from sH hydrate were observed, the methane
peak in the medium cage being at 2901.5 and at 2911.0 cm1
for the small cage, however, this study does not support such
an assignment.
Because of the complexity and low intensity of the LMGS
spectra, it is very difcult to quantify the LMGS in the hydrate
phase. Hence it is impossible to determine the cage occupancy
from Raman spectroscopy. Besides, an excess of free LMGS is
always present in the sample, making it more difcult to separate the signals from the hydrate phase. Full LMGS conversion
or single crystals may be required to eliminate the interference
of signal from the liquid. Raman spectra from the LMGS liquids are also shown in the gures. Generally the LMGS signals
arising from the hydrate phase are shifted slightly to higher
wavenumber. The OH stretching signals from sH hydrate are
comparable to those from ice or sI hydrate, with the most intense part of the signal at 3100 cm1 . Thus one has to be
rather careful if one wishes to use the OH signals from Raman
to identify hydrate structure (Schicks et al., 2005). Especially
for solid mixtures likely it is not possible. Raman spectra obtained for complex methane-containing hydrates will not offer
much information in terms of structure and cage occupancy especially when more than one guest molecule is present and can
only be used in conjunction with additional data. A combined
analysis with PXRD and/or NMR is therefore required to interpret the Raman spectra. From within the criteria of the tight
cageloose cage model, combined with the fairly large natural
line widths, one does not expect a great deal of discrimination
amongst the group of either tight or loose cages.
An extra experiment was performed to further verify the
arguments on peak assignment as presented above. Hydrate
was formed in a pressure cell with a quartz window for in situ
Raman analysis during hydrate formation and decomposition.
The design of the Raman cell is given elsewhere (Schicks and
Ripmeester, 2004). Hydrate was grown from ice powder with
TBME sprayed on top of the ice at 5 C and 3.5 MPa. The
hydrate was allowed to grow for 2 h before the temperature
was ramping up to 5 C for additional reaction (20 min). A
spectrum was recorded every minute by accumulating 5 scans

3938

R. Susilo et al. / Chemical Engineering Science 62 (2007) 3930 3939


t = 0 min at -5C
to = 120 min at -5C
t = (to+1) min at -2C

Intensity

2918

t = (to+2) min at 0.5C


t = (to+3) min at 2C
t = (to+4) min at 3.5C
t = (to+5) min at 5C
t = (to+6) min at 5C
t = (to+7) min at 5C
t = (to+8) min at 5C
t = (to+9) min at 5C
t = (to+10) minat 5C

2913

t = (to+15) minat 5C
t = (to+20) minat 5C

2890

2900

2910

2920

2930

Wavenumber (cm

2940

2950

-1)

Fig. 12. Raman spectra during hydrate formation from ice+TBME+methane


at 3.5 MPa.

td = 0 min at 8C

2918

td = 1 min at 10C

gree of conversion, gas content, and cage occupancies were obtained. The stable crystal structure is sH hydrate whenever there
is an excess of the large molecule guest substance (LMGS).
The presence of only n-heptane (as a non-hydrate former) results in a sI hydrate product. The methane content in sH hydrate increases in the following order: TBME<NH<MCH and
is opposite to the trend in the observed rate of hydrate growth.
Forming hydrate at a higher gas pressure increases the cage
occupancy. Hydrate conversion is limited when the amount of
LMGS used is less than the stoichiometric amount and when
the methane pressure (sH stability region) is low. Increasing the
pressure above the sI hydrate equilibrium line produces a mixture of sI and sH hydrates. The optimum gas storage capacity
of sH hydrate varies with the LMGS used and is approximately
2040% less than that of sI methane hydrate. The hydrate structure, degree of conversion, and gas content in the hydrate phase
obtained from the different techniques are all consistent with
each other. Methane and LMGS signals could be identied with
Raman and NMR spectroscopy, however, the hydrate composition can only be acquired through NMR spectroscopy. Finally,
some ambiguities in Raman results for sH hydrate reported in
the literature have been claried.

td = 2 min at 10C

Acknowledgments

td = 3 min at 10C

Intensity

td = 4 min at 10C

The nancial support from Natural Sciences and Engineering


Research Council of Canada (NSERC) is greatly appreciated.
Robin Susilo gratefully acknowledges nancial support from
Canada Graduate Scholarship (CGS).

2913

References

2890

2900

2910

2920

2930

2940

2950

Wavenumber (cm-1)
Fig. 13. Raman spectra during (CH4 + TBME) hydrate decomposition at
3.5 MPa.

(5 s), as shown in Fig. 12. The LMGS signals were invisible


due to the few scans acquired and very low signal intensity.
Initially, only methane gas was seen at 2918 cm1 and then
another signal grew at 2913 cm1 which corresponded to
hydrate formation. No other signal was observed in this spectral
region. The hydrate was then decomposed by increasing the
temperature above the phase equilibrium line (10 C). Fig. 13
shows the disappearance of the signal at 2913 cm1 leaving
only the gas phase. This further supports the fact that methane
molecules in small and medium cages of sH hydrate is not
distinguishable using Raman spectroscopy.
4. Conclusions
Structure I and structure H gas hydrate samples were synthesized, characterized and analyzed with PXRD, DSC, NMR, and
Raman spectroscopy. Hydrate structures, lattice constants, de-

Anderson, G.K., 2004. Enthalpy of dissociation and hydration number of


methane hydrate from the Clapeyron equation. Journal of Chemical
Thermodynamics 36, 11191127.
Chou, I.-M., Sharma, A., Burruss, R.C., Shu, J., Mao, H.-K., Hemley, R.J.,
Goncharov, A.F., Stern, L.A., Kirby, S.H., 2000. Transformation in methane
hydrate. Proceedings of the National Academy of Sciences 97 (25),
1348413487.
Collins, M.J., Ratcliffe, C.I., Ripmeester, J.A., 1990. Nuclear magnetic
resonance studies of guest species in clathrate hydrates: line-shape
anisotropies, chemical shifts, and the determination of cage occupancy
ratios and hydration numbers. Journal of Physical Chemistry 94, 157162.
Dalmazzone, D., Kharrat, M., Lachet, V., Fouconnier, B., Clausse, D.,
2002. DSC and PVT measurementsmethane and trichlorouoromethane
hydrate dissociation equilibria. Journal of Thermal Analysis and
Calorimetry 70, 493505.
Davidson, D.W., 1972. Clathrate hydrates. In: Franks, F. (Ed.), Water. A
Comprehensive Treatise, vol. 2. Plenum Press, New York.
Englezos, P., Lee, J.-D., 2005. Gas hydrates: a cleaner source of energy
and opportunity for innovative technologies. Korean Journal of Chemical
Engineering 22, 671681.
Giavarini, C., Maccioni, F., Santarelli, M.L., 2006. Modulated DSC for
gas hydrates analysis. Journal of Thermal Analysis and Calorimetry 84,
419424.
Gulluru, D.B., Devlin, J.P., 2006. Rates and mechanisms of conversion of ice
nanocrystals to ether clathrate hydrates: guest-molecule catalytic effects at
120 K. Journal of Physical Chemistry A 110, 19011906.
Khokhar, A.A., Gudmundsson, J.S., Sloan, E.D., 1998. Gas storage in structure
H hydrates. Fluid Phase Equilibria 150, 383392.
Kini, R., Dec, S.F., Sloan Jr., E.D., 2004. Methane + propane structure
II hydrate formation kinetics. Journal of Physical Chemistry A 108,
95509556.

R. Susilo et al. / Chemical Engineering Science 62 (2007) 3930 3939


Komai, T., Kang, S.-P., Yoon, J.-H., Yamamoto, Y., Kawamura, T., Ohtake,
M., 2004. In situ Raman spectroscopy investigation of the dissociation
of methane hydrate at temperatures just below the ice point. Journal of
Physical Chemistry B 108, 80628068.
Lee, J.-D., Susilo, R., Englezos, P., 2005. Kinetics of structure H gas hydrate.
Energy and Fuels 19, 10081015.
Linstrom, P.J., Mallard, W.G. (Eds.), 2005. NIST Chemistry WebBook,
NIST Standard Reference Database Number 69, June, National Institute
of Standards and Technology, Gaithersburg, MD, 20899 http://webbook.
nist.gov.
Mehta, A.P., Sloan, E.D., 1996. Improved thermodynamics parameters
for prediction of structure H hydrate equilibria. A.I.Ch.E. Journal 42,
20362046.
Mori, Y.H., 2003. Recent advances in hydrate-based technologies for natural
gas storagea review. Journal of Chemical Industry and Engineering
(China) 54 (Suppl.), 117.
Parlour, P.L., Dalmazzone, C., Herzhaft, B., Rousseau, L., Mathonat, C.,
2004. Characterization of gas hydrates formation using a new high pressure
MICRO-DSC. Journal of Thermal Analysis and Calorimetry 78, 165172.
Pietrass, T., Gaede, H.C., Bifone, A., Pines, A., Ripmeester, J.A., 1995.
Monitoring xenon clathrate hydrate formation on ice surfaces with optically
enhanced 129Xe NMR. Journal of the American Chemical Society 117,
75207525.
Ripmeester, J.A., Ratcliffe, C.I., 1988. Low-temperature cross-polarization/
magic angle spinning carbon-13 NMR of solid methane hydrates: structure,
cage, occupancy, and hydration number. Journal of Physical Chemistry 92
(2), 337339.
Ripmeester, J.A., Ratcliffe, C.I., 1996. Solid state NMR spectroscopy. In:
Davies, J.E.D., Ripmeester, J. (Eds.), A Comprehensive Supramolecular
Chemistry, vol. 8, Physical Methods in Supramolecular Chemistry.
Pergamon, Elsevier, Oxford.
Ripmeester, J.A., Ratcliffe, C.I., 1999. On the contributions of NMR
spectroscopy to clathrate science. Journal of Structural Chemistry 40 (5),
654662.
Ripmeester, J.A., Tse, J.S., Ratcliffe, C.I., Powell, B.M., 1987. A new clathrate
hydrate structure. Nature 325, 135136.
Schicks, J.M., Ripmeester, J.A., 2004. The coexistence of two different
methane hydrate phases under moderate pressure and temperature
conditions: kinetic versus thermodynamic products. Angewandte Chemie
International Edition 43, 33103313.
Schicks, J.M., Erzinger, J., Ziemann, M.A., 2005. Raman spectra of gas
hydratesdifferences and analogies to ice 1 h and (gas saturated) water.
Spectrochimica Acta Part A 61, 23992403.
Seo, Y.-T., Lee, H., 2003. 13 C NMR analysis and gas uptake measurements
of pure and mixed gas hydrates: development of natural gas transport and
storage method using gas hydrate. Korean Journal of Chemical Engineering
20, 10851091.
Sloan, E.D., 2003. Clathrate hydrate measurements: microscopic, mesoscopic,
and macroscopic. Journal of Chemical Thermodynamics 35, 4153.
Subramanian, S., Kini, R.A., Dec, S.F., Sloan Jr., E.D., 2000. Evidence of
structure II hydrate formation from methaneethane mixtures. Chemical
Engineering Science 55, 19811999.

3939

Sum, A.K., Burruss, R.C., Sloan, E.D., 1997. Measurement of clathrate


hydrates via Raman spectroscopy. Journal of Physical Chemistry B 101,
73717377.
Sun, Q., Duan, T.-Y., Zheng, H.-F., Ji, J.-Q., Wu, X.-Y., 2005. Phase
transformation of methane hydrate under high pressure. Journal of
Chemical Physics 122 (2), 024174.
Susilo, R., Lee, J.D., Englezos, P., 2005. Liquidliquid equilibrium data
of water with neohexane, methylcyclohexane tert-butyl methyl ether nheptane and vaporliquidliquid equilibrium with methane. Fluid Phase
Equilibria 231 (1), 2026.
Susilo, R., Moudrakovski, I.L., Englezos, P., Ripmeester, J.A., 2006. Hydrate
kinetics study in the presence of nonaqueous liquid by NMR spectroscopy
and imaging. Journal of Physical Chemistry B 110, 2580325809.
Tsuji, H., Kobayashi, T., Ohmura, R., Mori, Y.H., 2005. Hydrate formation by
water spraying in a methane + ethane + propane gas mixture: an attempt
at promoting hydrate formation utilizing LMGS for sH hydrates. Energy
and Fuels 19, 869876.
Tulk, C.A., Ripmeester, J.A., Klug, D.D., 2000. The application of Raman
spectroscopy to the study of gas hydrates. Annals of the New York Academy
of Sciences 912, 859872.
Uchida, T., Takeya, S., Wilson, L.D., Tulk, C.A., Ripmeester, J.A., Nagao,
J., Ebinuma, T., Narita, H., 2003. Measurements of physical properties
of gas hydrates and in situ observations of formation and decomposition
processes via Raman spectroscopy and X-ray diffraction. Canadian Journal
of Physics 81, 351357.
Uchida, T., Ohmura, R., Ikeda, I.Y., Nagao, J., Takeya, S., Hori, A.,
2006. Phase equilibrium measurements and crystallographic analyses on
structure-H type gas hydrate formed from the CH4 CO2 neohexanewater
system. Journal of Physical Chemistry B 110, 45834588.
Udachin, K.A., Ratcliffe, C.I., Ripmeester, J.A., 2002a. Hydrate structure and
composition from single crystal X-ray diffraction: examples of structure
I, II, and H hydrates. Proceedings of the Fourth International Conference
on Gas Hydrates, vol. 2. pp. 604607.
Udachin, K.A., Ratcliffe, C.I., Ripmeester, J.A., 2002b. Single crystal
diffraction studies of structure I, II and H hydrates: structure, cage
occupancy and composition. Supramolecular Chemistry 2, 405408.
Wang, X., Schultz, A.J., Halpern, Y., 2002. Kinetics of methane hydrate
formation from polycrystalline deuterated ice. Journal of Physical
Chemistry A 106, 73047309.
Wilson, L.D., Tulk, C.A., Ripmeester, J.A., 2002. Instrumental techniques for
the investigation of methane hydrates: cross-calibrating NMR and Raman
spectroscopic data. Proceedings of the Fourth International Conference on
Gas Hydrates, vol. 2. pp. 614618.
Yeon, S.-H., Seol, J., Lee, H., 2006. Structure transition and swapping
pattern of clathrate hydrates driven by external guest molecules. Journal
of American Chemical Society 128 (38), 1238812389.
Yoon, J.-H., Yamamoto, Y., Komai, T., Haneda, H., 2003. Rigorous approach
to the prediction of the heat of dissociation of gas hydrates. Industrial and
Engineering Chemical Research 42, 11111114.
Yoon, J.-H., Kawamura, T., Yamamoto, Y., Komai, T., 2004. Transformation
of methane hydrate to carbon dioxide hydrate: in situ Raman spectroscopic
observations. Journal of Physical Chemistry A 108, 50575059.

Вам также может понравиться