You are on page 1of 8

Article

pubs.acs.org/JPCC

Molecular Dynamics Study on the Growth Mechanism of Methane


plus Tetrahydrofuran Mixed Hydrates
Jyun-Yi Wu, Li-Jen Chen, Yan-Ping Chen, and Shiang-Tai Lin*
Department of Chemical Engineering, National Taiwan University, Taipei 10617, Taiwan
*
S Supporting Information

ABSTRACT: Molecular dynamics (MD) simulations are


performed to analyze the dominating factors for the growth
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

of CH4 + THF mixed hydrates, and the results are compared


Downloaded via INDIAN INST OF TECH MADRAS on March 16, 2023 at 06:04:16 (UTC).

with the growth of single guest CH4 and THF hydrates. While
CH4 hydrate has a type I crystalline structure, the presence of
THF in the aqueous phase results in the growth the type II
structure hydrate. Compared to THF hydrates, the presence of
CH4 in the system enhances the dissociation temperature. The
growth rate of CH4 + THF mixed exhibits a maximum value at
about 290 K at 10 MPa. The growth rate is found to be
determined by two competing factors: (1) the adsorption of CH4 at the solid−liquid interface, which is enhanced with decreasing
temperature, and (2) the migration of THF to the proper site at the interface, which is enhanced with increasing temperature.
Above 290 K, which is about 10 K higher than the dissociation temperature of pure THF hydrate, the growth of cage can proceed
only when a sufficient amount of CH4 is adsorbed at the interface. The growth rate is dominated by the uptake of CH4 at the
interface, as in the case of pure CH4 hydrate. Below 290 K, the growth is not much affected by the presence of CH4. Instead, the
growth rate is determined by the rearrangement of THF molecules at the interface, as in the case of pure THF hydrate.

1. INTRODUCTION as much as 18 K12,14 when there is 5.56 mol % of THF in the


Clathrate hydrates are a class of crystalline solids consisting of aqueous phase. In fact, the presence of THF results in a different
small guest molecules (such as methane or CO2) encapsulated in crystalline structure. Pure methane hydrates are of type I
the cage spaces formed by hydrogen-bonded water molecules. structure, where methane molecules stay in both the small
Naturally occurring clathrate hydrates, mostly methane hydrates, (pentagonal dodecahedron, 512) and large (hexagonal truncated
are found in permafrost region and sea floors. The amount of trapezohedron, 51262) cages created by the hydrogen-bonded
methane trapped in the form of hydrate is so abundant such that water molecules. The presence of THF would result in the
it is considered as a potential source of energy.1−4 Furthermore, formation of type II structure, with methane molecules
due to the characteristic of high gas density at relatively high occupying the small 512 cages and the THF occupying the larger
temperature and low pressure, gas hydrates are also considered as (hexadecahedron, 51264) cages. Type I structure (sI) contains
a good means for transport and storage of large quantities of gas, two 512 and six 51262 cages, and type II structure (sII) consists of
such as natural gas,5,6 carbon dioxide,5 and hydrogens.7−11 sixteen 512 and eight 51264 cages. Although the occupancy of
The knowledge of thermodynamic stability of clathrate methane increases with pressure, THF almost always occupies all
hydrates and how their stability limit can be enhanced or the 51264 cages in the type II structure.16,17 Therefore, the loading
reduced is important for developing processes involving them. of methane in the methane + THF mixed hydrate is significantly
One possible means to change the dissociation temperature of less than that in pure methane hydrate.16
gas hydrates is to control the system pressure. For example, the While THF is a powerful thermodynamic promotor of gas
dissociation temperature of methane hydrate is 273 K at 25 atm hydrates, the growth rate of CH4 + THF hydrates is found to be
and can be enhanced with increasing pressure.12 An alternative slower than that of CH4 hydrates under the same subcooling
way to alter the stability of gas hydrate is to introduce additives to temperature.18 However, the fundamental reason for the slower
the aqueous solution. For example, alcohols (e.g., methanol, growth kinetics with the presence of THF is not clearly known.
ethanol) are known to be effective inhibitors that reduce the MD simulations have been used to provide molecular insights
dissociation temperature under constant pressure. On the other into many important issues of gas hydrates, including
hand, cyclic ethers (e.g., tetrahydropyran, cyclobutanone, nucleation,19−24 guest replacement,25−27 growth/melting mech-
methylcyclohexane) are promotors that increase the dissociation anism,28−31 and structure changes.32,33 It has been shown that
temperature of gas hydrates.13
Tetrahydrofuran (THF) is known to be a very effective Received: June 5, 2015
thermodynamic promoter of gas hydrates.10,12,14,15 The Revised: August 6, 2015
formation temperature of methane hydrate can be enhanced by Published: August 10, 2015

© 2015 American Chemical Society 19883 DOI: 10.1021/acs.jpcc.5b05393


J. Phys. Chem. C 2015, 119, 19883−19890
The Journal of Physical Chemistry C Article

Figure 1. Initial structure of the gas−liquid−hydrate three-phase model of CH4 + THF hydrate. This model corresponds to a liquid phase THF
concentration of xTHF
L = xTHF
H . The system contains 112 THF 416 CH4 and 1904 water molecules.

the solubility of CH4 in water, the transport of CH4 to the target temperature. After these pre-equilibration steps (time zero
hydrate/liquid interface, and the uptake of methane on the point in our analysis), long (up to 1.7 μs) NPT simulations were
interface are the key factors determining the growth rate of then performed for analysis.
methane hydrates.28 For pure THF hydrates, the transport of The leapfrog algorithm41 is used with a integration time step of
THF to the interface and the migration of interfacial THF are the 1 fs. The cutoff radius for both van der Waals and Coulomb
dominating growth factors.34,35 Recent work of Erfan-Niya et interactions is 0.95 nm. Long-range Coulomb interactions are
al.36 investigated the bulk properties of CH4 + THF mixed determined using the particle-mesh Ewald (PME)42 method.
hydrates. However, to the best of our knowledge, there has not The Nosé−Hoover thermostat43 with tau_t = 1 ps is used for
been any report on the kinetics and growth mechanism of CH4 + temperature control and the Parrinello−Rahman44 with tau_P =
THF mixed hydrates. Therefore, the aim of the present work is to 10 ps for the pressure control. Anisotropic pressure control (one
calculate the growth rates and analyze the growth mechanism of for the growing direction and one for the other two directions) is
CH4 + THF hydrate by MD simulations. Our results help to adopted for the liquid−hydrate−liquid−gas three-phase model.
provide insights into the kinetic inhibition effect of THF during Isotropic pressure control is adopted for the simulation without
the growth of methane hydrates. hydrate phase (for the diffusivity and solubility tests).
2.3. Force Field. The TIP4P-Ice force field45 is choose in this
2. SIMULATION DETAILS study for H2O molecules. This force field is known to reproduce
the phase boundary between liquid water and ice, the density of
2.1. Molecular Model. The liquid−hydrate−liquid−gas (L- both liquid water and ice, the melting enthalpy, and the slope of
H-L-G) three-phase model, as shown in Figure 1, was created the coexistence curve. For CH4 molecules, the OPLS-AA
using Materials Studio37 for the simulation of dissociation model46 is employed. The geometric mean is used for the off-
condition and growth rate of CH4 + THF hydrate. The √2 × diagonal terms for the Lennard-Jones (LJ) potential between
√2 × 2 sII hydrate crystal with (100) crystal face has each of its water and CH4. For THF molecules, model 7 of Girard et al.47
large cages filled with one THF molecule and each of small cages was slightly modified (see Table 1) to better reproduce the
filled with one CH4 molecule (a total of 32 THF and 64 CH4).
The (100) crystal face was chosen because it was found the Table 1. Density and Heat of Vaporization of THF at 300 K
fastest growth face of THF hydrate in a previous study.35 For the and 1 atm Obtained Using Different Scaling Factors for the LJ
ease of future analysis, we define a unit hydrate layer (HL) as the 12−6 Potential Parameters of Girad
repeating unit in the hydrate phase (two HLs in the hydrate
phase in Figure 1). Each HL can be further divided into four density
equally sized sublayers (SL) along the z-direction.34 Before filling fσa fε a (kg/m3) Hvap (kJ/mol)
in guest molecules, the empty sII lattice was annealed by heating Girard 7a 1 1 904 34.32
and cooling with the position of oxygen atoms fixed. This allows modified Girard 7 1.003 0.933 880 31.92
for the water molecules to rotate and reconstruct the hydrogen exptl data 880b 31.803 ± 0.012c
a
bond network such that a structure of zero net dipole moment Factors for modifying the LJ 12−6 parameters from Girad’s work47
can be obtained. The liquid phase initially contains only water σnew
i = fσ × σGirad
i and εnew
i = fε × εGirad
i with i = C, H, or O. σGirad
C =
and THF (with the ratio of THF:water being the same as that in 0.385, σO = 0.350, σH = 0.190 (nm), εGirad
Girad Girad
C = 0.190, εGirad
O = 0.360,
the hydrate phase) and free of any CH4. It is expected that both and εGirad
H = 0.150 (kJ/mol). bAt 1 atm and 298.15 K.48 cAt 0.25 atm
49
the CH4 diffusion and the THF rearrangement at hydrate 300.80 K.
interface (which has been shown to be important at such high
THF concentration condition34) will affect the growth of CH4 +
THF hydrate in our simulation. density and heat of vaporization of pure THF liquid. The off-
2.2. Molecular Dynamics (MD) Simulation. The MD diagonal LJ potential parameters between THF and water were
simulations were performed using GROMACS 4.5.38−40 The taken from previous study.34 The solubility of THF in water and
initial structure was first energy minimized to eliminate any bad the dissociation temperature of THF hydrate from these settings
contacts. A 20 ps MD simulation was then conducted at 200 K are in good agreement with experiment. (Note that the solubility
under constant volume (NVT) to further relax any extra stress in of THF was underestimated and the dissociation temperature of
the system. A subsequent temperature rising process was then THF hydrate was slightly overestimated if the geometric mean
performed to increase the temperature the desired value at a rate was used to determine the off-diagonal LJ parameters.) It should
of 0.5 K/ps, followed by an additional 100 ps simulation at the be noted that no data of the mixed CH4 + THF hydrates were
19884 DOI: 10.1021/acs.jpcc.5b05393
J. Phys. Chem. C 2015, 119, 19883−19890
The Journal of Physical Chemistry C Article

used for the parameter optimization. The off-diagonal LJ


parameters are summarized in Table 2.

Table 2. Off-Diagonal LJ 12−6 Potential Parameters between


THF and Water
CTHF−OH2O OTHF−OH2O
force field ε (kcal/mol) σ (Å) ε (kcal/mol) σ (Å)
geometric mean 0.0944 3.496 0.1301 3.334
modified34 0.2547 3.430 0.1509 3.163

2.4. Growth/Melting Rate. The growth rate, r, of hydrate is Figure 2. Dissociation condition of CH4 hydrates (green triangles),
calculated from the number of water molecules added to the THF hydrates (red diamonds), and CH4 + THF hydrates (blue squares)
hydrate phase from the 100 to 300 ns of the NPT simulation. The from MD simulations and the experiment (lines).12,53,54
data of initial 100 ns was taken as the time needed for the system
to reach steady growth. Nonetheless, the promoting effect of THF for CH4 is correctly
number of water molecules increased THL predicted.
r= 3.2. Growth Rate. The growth rates of CH4 + THF mixed
200 ns WHL (1)
hydrate are determined at temperatures below the dissociation
where THL and WHL are the thickness and number of water condition (312.5 K) at 10 MPa. Figure 3 illustrates the time
molecules in one hydrate layer (HL). The values of these two evolution of the hydrate phase thickness (in terms of hydrate
properties depends on the structure of the hydrates. For sII layers, HL) from three independent sets of simulations at three
hydrates (THF and CH4 + THF), THL = 17.2 Å and WHL = 272. temperatures. In general, the thickness increases with time, but
For sI hydrates (CH4), THL = 11.9 Å and WHL = 184. For better there are ups and downs during the period of our growth
statistics, the growth rates are averaged from the results of three simulation. The cause of pauses and temporary dissociation will
independent runs (generated using different random seeds for be discussed in the next section. The slope of the curves in Figure
the initial velocity). The method for determining the number of 3d (average of the independent runs) gives the growth rates. The
water used in hydrate is the same as the previous study and data of initial 100 ns in our simulation are discarded (see eq 1) to
shown in Figures S1−S3 of the Supporting Information.34 avoid the effect from the initial relaxation processes. (There is a
clear change of slope in Figure 3d in the initial 5−10 ns as a result
3. RESULTS AND DISCUSSION of the diffusion of methane to the aqueous phase.) Figure 4
3.1. Validation of the Force Field. The thermodynamic summarizes the growth rate of CH4 + THF hydrate with different
properties such as the solubility of guests in water, dissociation subcooling temperatures at 10 MPa. The growth rate exhibits a
temperatures of hydrates, and the thermodynamic promotion maximum value at 290 K (a subcooling temperature of 22.5 K).
effect of THF for CH4 hydrates can serve as a validation of the Furthermore, the growth rate appears to be independent of
force field used in this study. The solubility is determined using temperature below 280 K.
NPT simulation for a two-phase (gas and liquid) model (as in ref Table 3 shows the growth rate of CH4, THF, and CH4 + THF
34). The solubility of methane in water is found to be 0.128 mol/ hydrate at different temperature and pressure. Note that under
kg water (or 434 water per CH4) at 10 MPa and 305 K, which is the same degree of subcooling (e.g., 7.5 K at 10 MPa) the growth
in good agreement with the experiment,50 0.111 mol/kg water rate of CH4 hydrates, 4.08 ± 0.20 mm/s (275 K), is 3 times
(or 500 water per CH4) at 10.22 MPa and 303.2 K. THF and greater than that of CH4 + THF hydrate, 1.20 ± 1.82 mm/s (305
water are completely miscible at ambient conditions. However, K). Such a kinetic inhibition effect of THF to CH4 hydrate was
as shown in Figure S4a, the use of geometric mean for the off- also observed in a previous experimental study.18 However, it is
diagonal LJ terms between THF and water (see Table 2) results noteworthy that the comparison of growth rate between CH4
in partially miscible two-liquid phases. This implies that the hydrate and CH4 + THF hydrate is not meaningful because they
interactions between THF and water are too weak. The modified are of different crystalline structures. A more sensible comparison
off-diagonal parameters (Table 2) increases the attractive is between THF hydrate and CH4 + THF hydrate, where it is
interactions between THF and water and thus results in a found that CH4, a thermodynamic promoter of THF hydrate, is a
completely miscible mixture (Figure S4b) as observed in kinetic inhibitor below the dissociation temperature of THF
experiment.51 hydrate and a kinetic promotor near and above the dissociation
The dissociation temperature (Td) of hydrates is determined temperature of THF hydrate.
by performing several NPT simulations with different temper- It is noteworthy that factors such as the size of the molecular
atures under the same pressure. If the temperature is above Td, model and the thickness of the aqueous layer can have an
the hydrate phase would melt and the potential energy would influence on the value of the growth rate. The strength of MD
increase. If the temperature is below Td, then the opposite would simulations is not to provide the quantitative value of growth rate
happen.27−30,34,52 Figure 2 compares the dissociation temper- but to provide insights, as will be discussed in the following
atures as a function of pressure for CH4, THF, and CH4 + THF section, into how the different factors (in our case the
hydrates from our simulation (open symbols) and experiment temperatures) affect the growth rate.
(lines). Both the dissociation temperatures of THF hydrate and 3.3. Growth Mechanism. In Figure 3, we observe that the
CH4 hydrate are well reproduced. The dissociation temperatures growth of CH4 + THF hydrate is not a steady process. For
of CH4 + THF hydrates are slightly overestimated by 5−10 K. example, in 10 MPa−275 K−run 3, the growth pauses at 20−60
19885 DOI: 10.1021/acs.jpcc.5b05393
J. Phys. Chem. C 2015, 119, 19883−19890
The Journal of Physical Chemistry C Article

Figure 3. Time evolution (growth) of CH4 + THF hydrate layers at 10 MPa and 275 K (a), 290 K (b), and 305 K (c). Shown in each figure are results of
three independent runs. The average values are compared in (d). The dissociation temperature at 10 MPa is 312.5 K.

interface are notably different. The difference can be understood


from the distribution of guest molecules at the interface. Figure 6
illustrates the cross section of the sublayer 1 (SL1) on the left-
and right-hand sides of the initial hydrate crystal at 30 ns and 1 μs.
At 30 ns (near the beginning of simulation), the sublayer 1 on
both sides are both solid−liquid interfaces. Both the THF and
CH4 molecules adsorbed on the small and large cage sites (see
Figure 6a,b). However, all the large cage sites are occupied by
THF molecules, and some of small cage sites are filled by CH4
molecules after the growth finishes, i.e., Figure 6c,d. The
rearrangement of guest molecules to the proper sites at the
interface is necessary for the successful growth of a SL.
To investigate the reason for intermittency in the growth
Figure 4. Growth rate of CH4 + THF hydrate as a function of process in more detail, the number of guest molecules (THF and
temperature at 10 MPa. CH4) absorbed to different sublayers (SL) during the course of
simulation is analyzed. Figure 7 shows the number of guest
Table 3. Growth Rate of CH4, THF, and CH4 + THF Hydrates molecules (THF and CH4) absorbed to different sublayers (SL)
from MD Simulations at 297.5 K during the course of simulation. It should be noted
system P (MPa) T (K) growth rate (mm/s) that each SL has four large cage (51264) sites and eight small cage
CH4 hydrate (512) sites (see Figure 6). As the growth process begins, the THF
10 260 3.56 molecules are adsorbed on both large and small cage sites, and
10 267.5 3.62 ± 0.61 thus the total number of THF quickly exceeds 4 (see SL1 before
THF hydrate 10 275 4.08 ± 0.20 200 ns (Figure 7a), SL2 from 200 to 400 ns (Figure 7b), and SL3
0.1 270 8.15 ± 1.57 from 400 to 700 ns (Figure 7c)). To proceed with the growth, the
CH4 + THF hydrate 10 262.5 7.37 ± 0.80 excess THF molecules, in particular those adsorbed on the small
10 262.5 3.91 ± 1.48 cage sites, must be desorbed. At high temperatures (e.g., above
10 270 3.13 ± 2.24 dissociation temperature of pure THF hydrate, 276 K at 10
10 275 2.83 ± 3.06 MPa), the THF molecules can escape from the interfacial cage
10 280 2.92 ± 0.81 sites more easily. As can be seen in SL3 from 400 to 700 ns, the
10 285 2.93 ± 2.02 number of THF molecules in both large and small cage sites
10 290 6.10 ± 2.89 changes quickly. Notice that the growth of SL3 (from 390 to 820
10 297.5 2.28 ± 2.10 ns) is slower than that of SL1 (from 0 to 180 ns) and SL2 (from
10 305 1.20 ± 1.84 180 to 390 ns). One obvious reason is the number of CH4 in SL3
remained at a low value (fluctuating between 0 and 1 most of the
ns, 70−145 s, 180−220 ns, and 240−270 ns. The duration of time) compared to those in other SLs. Before 700 ns, the number
pauses may take as long time as 100 ns (e.g., 70−300 ns in 10 of CH4 molecules at the interface is always equal to or smaller
MPa−275 K−run 2 or 20−300 ns in 10 MPa−305 K−run 3). than 2 (Figure 8g). After 700 ns, the number reaches 4 or 5 and
These pauses result in a slower value and higher uncertainty in the growth of the SL quickly finishes. In other words, the
the calculated growth rate. Figure 5 illustrates an example of the adsorption of enough CH4 at the interface is important at this
growth of CH4 + THF hydrate at 10 M Pa and 297.5 K. After 1 μs temperature. This is sensible because without CH4 the THF
(Figure 5b) the thickness of the CH4 + THF hydrate increased to hydrate would not from as the simulation temperature is already
4 hydrate layers. The growth rates of the left interface and right above the dissociation temperature of THF hydrate. In other
19886 DOI: 10.1021/acs.jpcc.5b05393
J. Phys. Chem. C 2015, 119, 19883−19890
The Journal of Physical Chemistry C Article

Figure 5. Snapshots of CH4 + THF hydrate growth at 10 MPa and 297.5 K at 30 ns (a) and 1 μs (b). The location of the solid−liquid interface at time 0
ns is indicated by bold dashed lines. The blue dashed lines are used to illustrate the ID of sublayers (SL) for discussion in the main text.

Figure 6. Cross section of left and right (b, d) sublayer one (SL1) from CH4 + THF hydrate growth at 30 ns (a: left surface, b: right surface) and 1 μs (c:
left surface, d: right surface).

words, the occupancy of CH4 in the small cages is important to methane is enhanced.28 However, the rearrangement of THF at
stabilize the hydrate structure at high temperatures. the interface is easier at high temperatures because of enhanced
Figure 8 shows the situation at 262.5 K, which is below the molecular mobility.34 The different temperature dependence of
dissociation temperature of pure THF hydrate. In this case, the these two competing factors results in a maximum growth rate at
presence of CH4 molecules is not indispensable for the growth of a temperature (about 290 K) that is slightly higher than the
structure II hydrate. As can be seen in Figure 8e,f, the growth of dissociation temperature of pure THF hydrate (276 K). It is
the hydrate layer can proceed even without any CH4 at the reasonable to presume that between 276 and 290 K the growth is
interface. On the other hand, the rearrangement of THF at the dominated by the rearrangement of THF.
interface becomes more difficult because of the lowered mobility The temperature dependence of growth rate also affects the
of water and THF. (The result was shown in the previous study; occupancy of CH4 given in Table 4. At 10 MPa and 290 K the
the ideal liquid THF concentration in THF hydrate growth is CH4 occupancy of small cages reaches a minimum value of 0.269.
higher in higher temperature.34) For example, the growth of SL2 Above 290 K, the occupancy of CH4 must be higher because the
(from 80 to 150 ns) does not finish until all the THF molecules hydrate interface would not be stable otherwise. Below 290 K, the
leave the small cage sites. The rate-determining process thus low growth rate (limited by THF rearrangement) allowed
changes to the rearrangement of guest molecules at the interface enough time for more CH4 adsorbing on the interface. The
at low temperatures. occupancy of the large cages is close to unity in all cases. This is
It is interesting to note that the adsorption of CH4 at the because that the THF hydrates are not stable without the large
interface is favored at lower temperatures as the solubility of cages being occupied.34 The high concentration of THF results
19887 DOI: 10.1021/acs.jpcc.5b05393
J. Phys. Chem. C 2015, 119, 19883−19890
The Journal of Physical Chemistry C Article

Figure 7. Number of THF (a, b, and c for layers 1, 2, and 3, respectively) and CH4 (d, e, and f for layers 1, 2, and 3, respectively) molecules located at the
aqueous phase near the left-hand-side interface. The system is at 10 MPa and 297.5 K.

Figure 8. Number of THF (a, b, and c for layers 1, 2, and 3, respectively) and CH4 (d, e, and f for layers 1, 2, and 3, respectively) molecules located at the
aqueous phase near the left-hand-side interface. The system is at 10 MPa and 262.5 K.

in fast adsorption of THF to the large cage sites, and thus the 4. CONCLUSION
growth rate is dominated by slower processes, adsorption of CH4 The growth mechanism of methane and tetrahydrofuran
at high temperatures and THF desorption/migration at low clathrate hydrates was investigated by MD simulations. The
temperatures. Also shown in Table 4 is the occupancy reported force field used provides reasonable accuracy for a variety of
from NMR study at 4 MPa and 292 K.16 The occupancy from our relevant thermodynamic properties, including the melting point
of ice, dissociation temperature of THF, CH4, and CH4 + THF
simulation at 5 MPa and 290 K agrees quite well with the hydrates, and solubility of CH4 and THF in water. The kinetic
experiment, which again confirms the reliability of our inhibition effect of THF in CH4 hydrate was also observed. Our
simulation. simulations show that the growth of CH4 + THF mixed hydrate
19888 DOI: 10.1021/acs.jpcc.5b05393
J. Phys. Chem. C 2015, 119, 19883−19890
The Journal of Physical Chemistry C Article

Table 4. Occupancy of Guests in the Newly Grown CH4 + THF Hydrate Layers
512 51264
source P (MPa) T (K) CH4 THF CH4 THF CH4 + THF
16
NMR 4 292 0.3701 0.9989 0.9989
MD 5 290 0.4262 0 0.0227 0.9769 0.9996
10 262.5 0.4419 0 0 0.9694 0.9694
10 270 0.5900 0 0.0128 0.9872 1.0000
10 275 0.4175 0 0.0198 0.9802 1.0000
10 280 0.4312 0 0.0025 0.9662 0.9687
10 285 0.3654 0 0.0003 0.9997 1.0000
10 290 0.2937 0 0.0519 0.9481 1.0000
10 297.5 0.4224 0 0 1 1
10 305 0.6141 0 0 1 1

is dominated by the adsorption of CH4 to the growing interface (4) Kvenvolden, K. A. Gas Hydrates - Geological Perspective and
and the migration and rearrangement of THF at the interface. At Global Change. Rev. Geophys. 1993, 31, 173−187.
temperature higher than the dissociation temperature of pure (5) Chatti, I.; Delahaye, A.; Fournaison, L.; Petitet, J. P. Benefits and
THF hydrate, the presence of CH4 provides the necessary lattice Drawbacks of Clathrate Hydrates: A Review of Their Areas of Interest.
Energy Convers. Manage. 2005, 46, 1333−1343.
framework in order to stabilize the crystalline structure.
(6) Gbaruko, B. C.; Igwe, J. C.; Gbaruko, P. N.; Nwokeoma, R. C. Gas
Therefore, the growth rate is dominated by the rate of CH4 Hydrates and Clathrates: Flow Assurance, Environmental and
accumulation at interface at high temperatures. At temperature Economic Perspectives and the Nigerian Liquified Natural Gas Project.
below the melting point of THF hydrate, the growth can proceed J. Pet. Sci. Eng. 2007, 56, 192−198.
without the aid of CH4. Under such circumstances, the growth (7) Chapoy, A.; Anderson, R.; Tohidi, B. Low-Pressure Molecular
rate is dominated by the escape of THF from the incorrect sites Hydrogen Storage in Semi-Clathrate Hydrates of Quaternary
on the interface, which is also the dominating factor for the Ammonium Compounds. J. Am. Chem. Soc. 2007, 129, 746−747.
growth of pure THF hydrate observed previously.34 The (8) Kim, D.-Y.; Park, Y.; Lee, H. Tuning Clathrate Hydrates:
different temperature dependence of these two phenomena Application to Hydrogen Storage. Catal. Today 2007, 120, 257−261.
results in a competing effect. The adsorption of CH4 is enhanced (9) Prasad, P. S. R.; Sowjanya, Y.; Shiva Prasad, K. Micro-Raman
Investigations of Mixed Gas Hydrates. Vib. Spectrosc. 2009, 50, 319−
as the temperature is decreased, but the migration of THF is
323.
enhanced with increasing temperature. As a result, the growth (10) Sugahara, T.; Haag, J. C.; Prasad, P. S. R.; Warntjes, A. A.; Sloan, E.
rate of CH4 + THF hydrate exhibits a maximum value at around D.; Sum, A. K.; Koh, C. A. Increasing Hydrogen Storage Capacity Using
290 K at 10 MPa. Tetrahydrofuran. J. Am. Chem. Soc. 2009, 131, 14616−14617.


*
ASSOCIATED CONTENT
S Supporting Information
(11) Ogata, K.; Tsuda, T.; Amano, S.; Hashimoto, S.; Sugahara, T.;
Ohgaki, K. Hydrogen Storage in Trimethylamine Hydrate: Thermody-
namic Stability and Hydrogen Storage Capacity of Hydrogen Plus
Trimethylamine Mixed Semi-Clathrate Hydrate. Chem. Eng. Sci. 2010,
The Supporting Information is available free of charge on the 65, 1616−1620.
ACS Publications website at DOI: 10.1021/acs.jpcc.5b05393. (12) de Deugd, R. M.; Jager, M. D.; de Swaan Arons, J. Mixed Hydrates
Procedure of identifying the water structures in hydrates of Methane and Water-Soluble Hydrocarbons Modeling of Empirical
Results. AIChE J. 2001, 47, 693−704.
and the results of THF solubility in water (PDF)
(13) Mooijer-van den Heuvel, M. M.; Peters, C. J.; de Swaan Arons, J.

■ AUTHOR INFORMATION
Corresponding Author
Influence of Water-Insoluble Organic Components on the Gas Hydrate
Equilibrium Conditions of Methane. Fluid Phase Equilib. 2000, 172, 73−
91.
(14) Chari, V. D.; Sharma, D. V. S. G. K.; Prasad, P. S. R. Methane
*E-mail stlin@ntu.edu.tw (S.-T.L.). Hydrate Phase Stability with Lower Mole Fractions of Tetrahydrofuran
Notes (Thf) and Tert-Butylamine (T-Bunh2). Fluid Phase Equilib. 2012, 315,
The authors declare no competing financial interest. 126−130.


(15) Strobel, T. A.; Koh, C. A.; Sloan, E. D. Thermodynamic
Predictions of Various Tetrahydrofuran and Hydrogen Clathrate
ACKNOWLEDGMENTS Hydrates. Fluid Phase Equilib. 2009, 280, 61−67.
The authors are grateful to the financial support of this research (16) Seo, Y. T.; Lee, H. C-13 Nmr Analysis and Gas Uptake
from the Ministry of Economic Affairs (103-5226904000-03-03) Measurements of Pure and Mixed Gas Hydrates: Development of
and the Ministry of Science and Technology (MOST 103-3113- Natural Gas Transport and Storage Method Using Gas Hydrate. Korean
M-002-006) of Taiwan. J. Chem. Eng. 2003, 20, 1085−1091.


(17) Yoon, J.-H. A Theoretical Prediction of Cage Occupancy and
Heat of Dissociation of Thf-Ch4 Hydrate. Korean J. Chem. Eng. 2012, 29,
REFERENCES 1670−1673.
(1) Dickens, G. R.; Paull, C. K.; Wallace, P. Direct Measurement of in (18) Lim, S.-H.; Riffat, S. B.; Park, S.-S.; Oh, S.-J.; Chun, W.; Kim, N.-J.
Situ Methane Quantities in a Large Gas-Hydrate Reservoir. Nature Enhancement of Methane Hydrate Formation Using a Mixture of
1997, 385, 426−428. Tetrahydrofuran and Oxidized Multi-Wall Carbon Nanotubes. Int. J.
(2) Gornitz, V.; Fung, I. Potential Distribution of Methane Hydrates in Energy Res. 2014, 38, 374−379.
the Worlds Oceans. Global Biogeochem. Cycles 1994, 8, 335−347. (19) Jacobson, L. C.; Hujo, W.; Molinero, V. Amorphous Precursors in
(3) Haq, B. U. Natural Gas Deposits - Methane in the Deep Blue Sea. the Nucleation of Clathrate Hydrates. J. Am. Chem. Soc. 2010, 132,
Science 1999, 285, 543−544. 11806−11811.

19889 DOI: 10.1021/acs.jpcc.5b05393


J. Phys. Chem. C 2015, 119, 19883−19890
The Journal of Physical Chemistry C Article

(20) Jacobson, L. C.; Molinero, V. Can Amorphous Nuclei Grow (41) Hockney, R. W.; Goel, S. P.; Eastwood, J. W. Quiet High-
Crystalline Clathrates? The Size and Crystallinity of Critical Clathrate Resolution Computer Models of a Plasma. J. Comput. Phys. 1974, 14,
Nuclei. J. Am. Chem. Soc. 2011, 133, 6458−6463. 148−158.
(21) Walsh, M. R.; Beckham, G. T.; Koh, C. A.; Sloan, E. D.; Wu, D. T.; (42) Darden, T.; York, D.; Pedersen, L. Particle Mesh Ewald - an
Sum, A. K. Methane Hydrate Nucleation Rates from Molecular N.Log(N) Method for Ewald Sums in Large Systems. J. Chem. Phys.
Dynamics Simulations: Effects of Aqueous Methane Concentration, 1993, 98, 10089−10092.
Interfacial Curvature, and System Size. J. Phys. Chem. C 2011, 115, (43) Nose, S. A Unified Formulation of the Constant Temperature
21241−21248. Molecular-Dynamics Methods. J. Chem. Phys. 1984, 81, 511−519.
(22) Walsh, M. R.; Koh, C. A.; Sloan, E. D.; Sum, A. K.; Wu, D. T. (44) Parrinello, M.; Rahman, A. Polymorphic Transitions in Single-
Microsecond Simulations of Spontaneous Methane Hydrate Nucleation Crystals - a New Molecular-Dynamics Method. J. Appl. Phys. 1981, 52,
7182−7190.
and Growth. Science 2009, 326, 1095−1098.
(23) Jacobson, L. C.; Hujo, W.; Molinero, V. Nucleation Pathways of (45) Abascal, J. L. F.; Sanz, E.; Fernandez, R. G.; Vega, C. A Potential
Model for the Study of Ices and Amorphous Water: Tip4p/Ice. J. Chem.
Clathrate Hydrates: Effect of Guest Size and Solubility. J. Phys. Chem. B
Phys. 2005, 122, 234511.
2010, 114, 13796−13807. (46) Jorgensen, W. L.; Maxwell, D. S.; TiradoRives, J. Development
(24) Jacobson, L. C.; Matsumoto, M.; Molinero, V. Order Parameters and Testing of the Opls All-Atom Force Field on Conformational
for the Multistep Crystallization of Clathrate Hydrates. J. Chem. Phys. Energetics and Properties of Organic Liquids. J. Am. Chem. Soc. 1996,
2011, 135, 074501. 118, 11225−11236.
(25) Geng, C.-Y.; Wen, H.; Zhou, H. Molecular Simulation of the (47) Girard, S.; Muller-Plathe, F. Molecular Dynamics Simulation of
Potential of Methane Reoccupation During the Replacement of Liquid Tetrahydrofuran: On the Uniqueness of Force Fields. Mol. Phys.
Methane Hydrate by Co2. J. Phys. Chem. A 2009, 113, 5463−5469. 2003, 101, 779−787.
(26) Qi, Y.; Ota, M.; Zhang, H. Molecular Dynamics Simulation of (48) Carvajal, C.; Tolle, K. J.; Smid, J.; Szwarc, M. Studies of Solvation
Replacement of Ch4 in Hydrate with Co2. Energy Convers. Manage. Phenomena of Ions and Ion Paris in Dimethoxyethane and
2011, 52, 2682−2687. Tetrahydrofuran. J. Am. Chem. Soc. 1965, 87, 5548−5553.
(27) Tung, Y.-T.; Chen, L.-J.; Chen, Y.-P.; Lin, S.-T. In Situ Methane (49) Hossenlopp, I. A.; Scott, D. W. Vapor Heat-Capacities and
Recovery and Carbon Dioxide Sequestration in Methane Hydrates: A Enthalpies of Vaporization of 6 Organic-Compounds. J. Chem.
Molecular Dynamics Simulation Study. J. Phys. Chem. B 2011, 115, Thermodyn. 1981, 13, 405−414.
15295−15302. (50) Wang, L. K.; Chen, G. J.; Han, G. H.; Guo, X. Q.; Guo, T. M.
(28) Tung, Y.-T.; Chen, L.-J.; Chen, Y.-P.; Lin, S.-T. The Growth of Experimental Study on the Solubility of Natural Gas Components in
Structure I Methane Hydrate from Molecular Dynamics Simulations. J. Water with or without Hydrate Inhibitor. Fluid Phase Equilib. 2003, 207,
Phys. Chem. B 2010, 114, 10804−10813. 143−154.
(29) Tung, Y.-T.; Chen, L.-J.; Chen, Y.-P.; Lin, S.-T. Growth of (51) Jones, C. Y.; Zhang, J. S.; Lee, J. W. Isotope Effect on Eutectic and
Structure I Carbon Dioxide Hydrate from Molecular Dynamics Hydrate Melting Temperatures in the Water-Thf System. J. Thermodyn.
Simulations. J. Phys. Chem. C 2011, 115, 7504−7515. 2010, 2010, 1−6.
(30) Tung, Y.-T.; Chen, L.-J.; Chen, Y.-P.; Lin, S.-T. Molecular (52) García Fernández, R.; Abascal, J. L. F.; Vega, C. The Melting Point
Dynamics Study on the Growth of Structure I Methane Hydrate in of Ice Ih for Common Water Models Calculated from Direct
Aqueous Solution of Sodium Chloride. J. Phys. Chem. B 2012, 116, Coexistence of the Solid-Liquid Interface. J. Chem. Phys. 2006, 124,
144506.
14115−14125.
(53) Gough, S. R.; Davidson, D. W. Composition of Tetrahydrofuran
(31) Zhang, J.; Hawtin, R. W.; Yang, Y.; Nakagava, E.; Rivero, M.; Choi,
Hydrate and the Effect of Pressure on the Decomposition. Can. J. Chem.
S. K.; Rodger, P. M. Molecular Dynamics Study of Methane Hydrate
1971, 49, 2691−2699.
Formation at a Water/Methane Interface. J. Phys. Chem. B 2008, 112, (54) Sloan, E. D.; Koh, C. A. Clathrate Hydrates of Natural Gases, 3rd
10608−10618. ed.; CRC Press: Boca Raton, FL, 2008.
(32) Jacobson, L. C.; Matsumoto, M.; Molinero, V. Order Parameters
for the Multistep Crystallization of Clathrate Hydrates. J. Chem. Phys.
2011, 135, 074501−1−074501−7.
(33) Nguyen, A. H.; Jacobson, L. C.; Molinero, V. Structure of the
Clathrate/Solution Interface and Mechanism of Cross-Nucleation of
Clathrate Hydrates. J. Phys. Chem. C 2012, 116, 19828−19838.
(34) Wu, J.-Y.; Chen, L.-J.; Chen, Y.-P.; Lin, S.-T. Molecular Dynamics
Study on the Equilibrium and Kinetic Properties of Tetrahydrofuran
Clathrate Hydrates. J. Phys. Chem. C 2015, 119, 1400−1409.
(35) Nada, H. Anisotropy in Growth Kinetics of Tetrahydrofuran
Clathrate Hydrate: A Molecular Dynamics Study. J. Phys. Chem. B 2009,
113, 4790−4798.
(36) Erfan-Niya, H.; Modarress, H.; Zaminpayma, E. Computational
Study on the Structure Ii Clathrate Hydrate of Methane and Large Guest
Molecules. J. Inclusion Phenom. Mol. Recognit. Chem. 2011, 70, 227−239.
(37) Materials Studio Modeling Environment; Accelrys Software Inc.:
San Diego, 2007.
(38) Pronk, S.; et al. Gromacs 4.5: A High-Throughput and Highly
Parallel Open Source Molecular Simulation Toolkit. Bioinformatics
2013, 29, 845−854.
(39) Berendsen, H. J. C.; Vanderspoel, D.; Vandrunen, R. Gromacs - a
Message-Passing Parallel Molecular-Dynamics Implementation. Com-
put. Phys. Commun. 1995, 91, 43−56.
(40) Van der Spoel, D.; Lindahl, E.; Hess, B.; Groenhof, G.; Mark, A.
E.; Berendsen, H. J. C. Gromacs: Fast, Flexible, and Free. J. Comput.
Chem. 2005, 26, 1701−1718.

19890 DOI: 10.1021/acs.jpcc.5b05393


J. Phys. Chem. C 2015, 119, 19883−19890

You might also like