Abstract
There is interest in the role of ammonia on Saturn’s moons Titan and Enceladus as the presence of water, methane, and ammonia under temperature and pressure conditions of the surface and interior make these moons rich environments for the study of phases formed by these materials. Ammonia is known to form solid hemi-, mono-, and dihydrate crystal phases under conditions consistent with the surface of Titan and Enceladus, but has also been assigned a role as water-ice antifreeze and methane hydrate inhibitor which is thought to contribute to the outgassing of methane clathrate hydrates into these moons’ atmospheres. Here we show, through direct synthesis from solution and vapor deposition experiments under conditions consistent with extraterrestrial planetary atmospheres, that ammonia forms clathrate hydrates and participates synergistically in clathrate hydrate formation in the presence of methane gas at low temperatures. The binary structure II tetrahydrofuran + ammonia, structure I ammonia, and binary structure I ammonia + methane clathrate hydrate phases synthesized have been characterized by X-ray diffraction, molecular dynamics simulation, and Raman spectroscopy methods.
Keywords: ice, single crystal X-ray diffraction, hydrogen bonding, hydrate inhibitors, ethane
Ammonia has long been seen as a key species in extraterrestrial space, both interstellar and on outer planets, moons, and comets and the interplay of ammonia, methane, and water has been the subject of a considerable number of studies and speculation (1–8). The main role assigned to ammonia has been that of an antifreeze for ice and clathrate hydrate formation, modifying the stability region of the solid ice and methane clathrate hydrate phases as a thermodynamic inhibitor (2, 5, 9). However, ammonia is a methane-sized molecule, thus based on size alone it has the potential for being a suitable guest for clathrate hydrate cages. Issues that may have prevented ammonia from being considered as a suitable clathrate guest molecule include the notion that guest species need to be hydrophobic in order to be incorporated into clathrates, and the observation of a number of stoichiometric nonclathrate phases of ammonia and water obtained upon cooling aqueous ammonia solutions (2). Previous experimental work on the water-ammonia and water-methane-ammonia systems had not shown evidence for the enclathration of ammonia (5, 10–19). Close inspection, however, shows that the low pressure ammonia dihydrate (18) and the high pressure phase II of ammonia monohydrate (19) have structural features in common with canonical clathrate and semiclathrate structures.
Recent structural analysis and molecular simulations have shown that some guest molecules which form strong hydrogen bonds with the water framework of the clathrate hydrate lattice may nonetheless produce stable phases (20–23). It is therefore reasonable to consider that ammonia has potential as a clathrate guest molecule. Furthermore, previous experimental and computational studies have shown that in the gas phase, water forms dodecahedral cages around ammonia and ammonium ions. These structures may be indicative of a geometric tendency towards clathrate hydrate structures for the ammonia and ammonium molecules (24–27).
Results
As a first step in this study we show that in the presence of other clathrate hydrate forming substances, ammonia indeed can be incorporated in cages within the clathrate lattice. Incorporation of ammonia in a clathrate hydrate phase was accomplished by freezing a solution of the well known clathrate hydrate former tetrahydrofuran (THF) and approximately 5% aqueous ammonia with an approximate THF:water mole ratio of 1∶17 at -10 °C. Crystals formed were harvested after a few days and those suitable for diffraction were identified and mounted on a diffractometer at low temperatures. The above experimental conditions can be compared with those in the study of Dong et al. (15, 16) who used THF:water ratios of about 1∶63 with up to 5% ammonia to study methane hydrate formation in the presence of ammonia. In their experiments, Dong et al. (15, 16) did not use sufficient THF to lead to the direct formation of the binary THF + NH3 clathrate hydrate and ammonia hydrate formation was not observed.
Structural data were recorded at 100 K (see Materials and Methods). The structure was shown to be the usual cubic structure II (sII) clathrate hydrate (28), space group Fd-3m with a unit cell edge of 1.71413(6) nm. The THF molecules occupy the large cages, as expected, and 39% of the small cages are filled with ammonia molecules (Fig. 1). This measured small cage occupancy is likely to be a function of the starting composition. Whether small cages could be completely filled with a more ammonia-rich liquid is uncertain and should be studied separately. One of the water molecules in the small cages which encapsulate the ammonia guests has moved out of its normal position, 0.118 nm inward into the small cage thus forming H2O⋯H-NH2 hydrogen bonds (0.267–0.272 nm) with NH3 (as suggested by hydrogen positions on the ammonia guest). The formation of HOH⋯NH3 hydrogen bonding could also lead to this water displacement. This displacement breaks the hydrogen bond of this water molecule with another water from an adjacent large (or small) cage (O⋯O distance 0.393 nm), thus distorting both large and small cages. The equivalent O⋯O distance in the undistorted edges of the water polyhedra is 0.279 nm. It should be noted that the minimum heavy atom N⋯O distance in the clathrate hydrate is smaller than in the ammonia hydrates, where the more uniform local environment of the NH3 molecule results in N⋯O distances of 0.305 to 0.330 nm.
Fig. 1.
The large and small cages of the cubic structure II THF-ammonia binary clathrate hydrate from single crystal X-ray diffraction. The ammonia guest has moved a water molecule out of its normal position by pulling it into the small cavity by forming a H2N-H⋯OH2 or H3N⋯HOH hydrogen bond.
Because structural evidence for the mixed THF + NH3 sII hydrate showed that it is possible to incorporate ammonia into clathrate cages, attempts were made to form clathrate hydrates with only ammonia as guest. Pure ammonia clathrate hydrate synthesis cannot be done by simply cooling aqueous ammonia solutions, as a variety of stoichiometric hydrates of ammonia are known to form preferentially (2, 11, 18, 19). Other ways of forming clathrate hydrates include vapor deposition of water at low temperatures to yield amorphous ice, followed by exposure of the ice to a pressure of guest gas and annealing (29), or vapor codeposition of water and the potential guest material at low temperatures, again, followed by annealing (23, 30, 31). It has been shown that below approximately 140 K ice surfaces are relatively inert unless strong hydrogen-bond donors or acceptors are present. For example, molecules normally hydrolyzed by water such as formaldehyde will not do so when in contact with ices at low temperature and will template clathrate hydrate lattices when amorphous ice is annealed (30).
Following the latter procedure, described in detail in Materials and Methods, vapor codeposits of ammonia and water were prepared at T < 20 K. The codeposited crystals were harvested at 77 K and kept in liquid nitrogen until appropriate measurements could be made. A portion of each amorphous product was mounted in a low temperature cell on a powder X-ray diffractometer and the samples annealed stepwise, increasing temperature from approximately 100 K. Fig. 2 shows the transformation of the amorphous deposit upon annealing. Fig. 2A shows the powder X-ray diffraction (PXRD) pattern for amorphous ice with traces of ices Ic and Ih at 100 K. At 140 K, the amorphous ice has largely transformed to ices Ic and Ih plus a number of other crystalline phases (the Bragg positions for Ic are omitted). These reflections continue to develop with increase of temperature to 150 K. The inset (Fig. 2B) shows a vertical expansion of the 150 K pattern. For each pattern the profiles are matched to those of known crystalline phases, including the stoichiometric ammonia hydrates (hemi-, mono-, and dihydrates), hexagonal ice Ih, and clathrate hydrates. Eight reflections (shown in Fig. 2) can be assigned to the cubic structure I (sI) ammonia clathrate hydrate and indexed to be consistent with sI clathrate hydrate, space group Pm-3n, with unit cell edge 1.1818(2) nm at 150 K.
Fig. 2.
PXRD patterns of NH3 and H2O co-deposition sample (A) recorded at 100, 140, and 150 K; (B) expansion of the pattern at 150 K. Lattice parameters (150 K) are: a = 0.4527(1) nm, b = 0.5586(1) nm, c = 0.9767(3) nm for ammonia monohydrate (space group: P212121); a = 0.4496(1) nm and c = 0.7339(1) nm for ice Ih (space group P63/mmc); a = 1.1818(2) nm for sI ammonia clathrate hydrate (space group Pm-3n); a = 0.8404(2) nm, b = 0.8441(3) nm, and c = 5.345(1) nm for ammonia hemihydrate (space group Pbnm); and a = 0.7125(1) nm for ammonia dihydrate (space group P213).
The interactions of ammonia with methane clathrate hydrate are also considered to be important in determining the phase equilibria on Titan and Enceladus. Several vapor deposits of methane, ammonia, and water were prepared as described in Materials and Methods and annealing of the vapor codeposits was followed by PXRD. Fig. 3A shows the PXRD pattern at 112 K, indicating the material to be mainly amorphous except for some crystalline Ih. At 143 K (Fig. 3A, second trace) most of the pattern for the amorphous deposit has transformed into ice Ic plus other crystalline phases that upon closer inspection suggested the presence of both cubic sI and sII clathrate hydrates. At 160 K, this pattern is well developed and a vertically expanded detail is given in Fig. 3B. Profile matching for each powder pattern shows that a good fit is obtained by considering the presence of ices Ic and Ih, as well as cubic sI [Pm-3n, a = 1.1847(8) nm] and cubic sII clathrate hydrates [Fd-3m, a = 1.7161(9) nm]. To observe both common cubic clathrate hydrate structures is surprising. Structure I is the usual clathrate that forms for methane at low pressures although metastable sII methane hydrate has been observed as well (30). A blank run for a vapor codeposit of water and methane but without ammonia did not reveal any clathrate hydrate under similar conditions of annealing (Fig. S1). Further annealing of the three-component codeposit at 180 K showed (Fig. 3A) that the sII clathrate hydrate pattern had largely disappeared, and reflections for the stoichiometric ammonia hydrates were absent. The transformation from mixed sI and sII hydrates to sI hydrate was confirmed by observing methane C-H symmetric stretch frequencies around 2,905 cm-1 by Raman spectroscopy (32) (Fig. S2). At low temperature, the deposited methane peak appeared at 2,908 cm-1, which is close to the Raman shift of dissolved methane in water (33). This peak splits into a doublet at 160 K; one peak at 2,903 cm-1 representing the methane in the large cages and the second at 2,914 cm-1 representing methane in the small cages. As temperature increases, the ratio of methane in the large cages to that of the guest in the small cages increases, which implies the transformation to sI phase (six large cages to two small cages) which has a greater proportion of large cages than sII (8 large cages to 16 small cages). Similar enhanced processes of CO2 clathrate hydrate formation on ice nanoparticles at low temperatures in the presence of hydrogen bonding ether guest molecules have been observed (34).
Fig. 3.
PXRD patterns of NH3, CH4, and H2O co-deposition sample (A) recorded at 112, 143, 160, and 180 K; (B) expansion of the pattern at 160 K. Vertical thin red lines and arrows indicate the peaks from sII hydrate and blue lines and arrows indicate the peaks from sI hydrate. Lattice parameter (160 K): a = 0.4491(3) nm and c = 0.7333(5) nm for Ih (space group P63/mmc), a = 0.6361(3) nm for Ic (space group Fd-3m), a = 1.1847(8) nm for sI hydrate (space group Pm-3n), and a = 1.7161(9) nm for sII hydrate (space group: Fd-3m).
To study molecular level structure and dynamics of ammonia containing clathrate hydrates, we performed molecular dynamics simulations of the THF (in large cages) + NH3 (in 39% of small cages) binary sII clathrate hydrate from 100 to 240 K. We also simulated the pure NH3 sI and sII clathrate hydrates and binary CH4 (large cages) + NH3 (small cages) sII clathrate hydrates for temperatures in the range of 100 to 240 K. The molecular dynamics simulation methodology is described in detail in Materials and Methods. For the duration of the simulations, all the clathrates are mechanically stable at simulation temperatures up to 200 K and did not show signs of decomposition and cage collapse.
In the molecular simulations of the sI and sII pure ammonia hydrate phases, we observe hydrogen bonding between the ammonia guest molecules and the water molecules in both the small and large clathrate hydrate cages. Cases where ammonia is the hydrogen donor (H2N-H⋯OH2) and hydrogen acceptor (H3N⋯H-OH) are both common but the latter is more frequent (Figs. S3 and S4). On the other hand, in the binary sII hydrate phases with THF or methane in the large cages, ammonia showed either no hydrogen bonding or only a small probability for hydrogen bond formation with water molecules of the small cages (see Figs. S5 and S6). From integrating the hydrogen bonding peak in the radial distribution functions (RDF), we calculated the probability of hydrogen bonding, p, in the pure NH3 sI, sII, and binary NH3 + CH4 sII clathrate hydrates for ammonia molecules and the results are given in Table 1. Peaks in the RDF with H2N-H⋯OH2 or H3N⋯H-OH distances less than 0.25 nm were assigned to hydrogen bonding. The areas underneath the hydrogen bond peaks in the RDF give a measure of the hydrogen bond probability (see Materials and Methods). The probability of hydrogen bonding increases with temperature, which shows that the ammonia-water hydrogen bonding in the clathrate hydrates is endothermic. The temperature dependence of the hydrogen bonding probability is evaluated using the van’t Hoff plot of ln K [where K = p/(1 - p)] as a function of 1/T (Fig. S7). The slopes of the linear plots are used to determine the enthalpies of hydrogen bond formation for ammonia in the different environments which are also given in Table 1. The formation of hydrogen bonds in the sII clathrate hydrate is less endothermic and the more facile formation of these bonds may destabilize the sII clathrate hydrate water lattice to a greater extent than the sI case.
Table 1.
The average percent (H3N⋯H-OH) hydrogen bonding configurations calculated from RDFs and the enthalpy of hydrogen bond formation for pure sI and sII NH3 hydrates and binary sII CH4 (large cages) + NH3 (small cages) hydrate at different temperatures
| Guests | 100 K | 120 K | 140 K | 160 K | 170 K | ΔHHB/kJ·mol-1 |
| NH3 (pure sI, large cages) | 4.6 | 12.9 | 34.5 | 54.4 | 68.9 | 7.66 |
| NH3 (pure sI, small cages) | 1.4 | 2.1 | 15.0 | 32.0 | 49.8 | 8.90 |
| NH3 (pure sII, large cages) | 7.8 | 17.1 | 28.3 | 51.5 | 61.8 | 5.85 |
| NH3 (pure sII, small cages) | 1.4 | 4.3 | 6.1 | 24.4 | 33.6 | 6.95 |
| NH3 (binary sII, small cages) | ∼0 | 0.6 | 1.3 | 1.8 | 2.9 | 5.09 |
Sample hydrogen bonded configurations of the ammonia guests in the large and small cages of the sI and sII clathrate hydrates are shown in Fig. 4. Hydrogen bond configurations with ammonia acting as the proton donor and the proton acceptor (often simultaneously) are observed. The ammonia guest remains mostly in the cage but in some configurations the ammonia molecule displaces one of the water molecules from a lattice site. The water molecules in the hexagonal faces of the large clathrate hydrate cages are less hindered and may also have less stable water-water hydrogen bonds which require less energy to break compared to the water molecules forming hydrogen bonds in pentagonal faces. Snapshots from the simulations show that the ammonia molecules preferentially hydrogen bond with the water molecules in the hexagonal faces. Ammonia-water hydrogen bonding severely distorts the small and large cage structures. At temperatures below 200 K, the hydrate phase maintains its mechanical stability without framework collapse despite the ammonia-water hydrogen bonds and the defects they induce in the hydrate lattice hydrogen-bonding network. As we only performed a molecular dynamics simulation and not a full free energy calculation comparing the thermodynamic stability of the clathrate hydrate phase to the products, we cannot comment on the thermodynamic stability of the ammonia clathrate hydrate phase. By mechanical stability we imply that the hydrate structure did not decompose under the pressure-temperature conditions of the simulation. As temperature and vibrational amplitudes of the water molecules in the hydrate lattice increase, these lattice defects lead to the breakdown of the water framework structure and instability of the hydrate phase.
Fig. 4.
Sample configurations of the NH3 guests hydrogen bonding with the water molecules of the large (L) and small (S) cages of the sI and sII clathrate hydrate (Top: sI, Bottom: sII cages). The snapshots of the cages were extracted from the periodic simulation cell. In cases where NH3 is incorporated into the water lattice, the displaced water molecule is also shown.
From Table 1, the ammonia molecules in the large sI and sII cages form hydrogen bonds more readily than in the corresponding small cages. In addition, the larger hydrogen bonding formation enthalpy in the sI hydrate compared to the sII hydrate, reveals that the water lattice in the sI phase is less susceptible to hydrogen bonding with ammonia and is preferred when the ammonia forms a clathrate phase with amorphous ice upon annealing.
In contrast to the pure NH3 hydrates, binary CH4 + NH3 sII hydrate shows a H3N⋯H-OH hydrogen bond probability of only approximately 3% even at 170 K. Incorporating methane in the cages of the binary clathrate hydrate apparently stabilizes the cages which encapsulate ammonia. The molecular dynamics simulation results are consistent within the framework of experimental knowledge of the ammonia hydrate system described above. Strong hydrogen bonding of ammonia with water in the solid phase is expected. The stability of the ammonia hydrate is dependent on the low temperature and/or stabilizing effect of methane help gas.
Discussion
The presence of both ammonia and methane give a clathrate hydrate that is much more stable than that of the pure ammonia hydrate (stability up to 180 K for the mixed clathrate as compared to 150 K for the pure NH3 clathrate). On the other hand, the presence of ammonia is critical in forming a clathrate in a temperature region where methane by itself does not form a clathrate hydrate. The activation of ice surfaces by ammonia at low temperatures is a well known phenomenon (35), so it may well be that synergistic behavior of ammonia and methane on the amorphous ice surface produce a reasonably stable clathrate: ammonia to initiate the reaction by introducing defects and methane to stabilize the clathrate lattice by limiting guest-host hydrogen bonding. Experiments with ethane and ammonia gave similar results (Fig. S8). From the work presented it is clear that ammonia plays an important role, so far unrealized, in forming solid clathrate phases at low temperatures in planetary environments. It may be difficult to quantify the effects of ammonia on a clathrate hydrate-forming environment as it is clear that some ammonia phases have time and temperature-limited regions of existence.
Recent spectrophotometric studies by Nelson and coworkers have shown variations in the surface reflectance on Titan which have been attributed to episodic cryovolcanic effusion events which bring NH3 from the interior of Titan to its surface in the form of a fog or frost (36). Most work on ammonia present in Titan focuses on ammonia in aqueous solution in the “magma” and its effect on methane clathrate hydrate inhibition and decomposition in that region (5). Our work shows that under temperature and pressure conditions similar to those found on Titan’s surface (5) (T ≈ 90 K and Patm ≈ 1.6 bar), considering the presence of methane in the atmosphere of Titan and availability of ice on its surface, it is possible that the extruded ammonia vapor forms binary clathrate hydrate phases with methane. Clathrate hydrate formation after ammonia vapor deposition in the presence of methane gas and annealing to about 100 K can provide a mechanism for the removal of ammonia and methane from Titan’s atmosphere. Ethane clathrate hydrate formation may also be catalyzed by the presence of ammonia.
Similar cryovolcanic activity is also seen in Saturn’s moon Enceladus where plumes of ammonia, methane, and water vapor are extruded from the South Pole (37). The sources of these methane extrusions in both moons have been related to clathrate hydrate dissociation (5, 38, 39).
The broader significance of the catalyzing effect of ammonia on methane hydrate formation under vapor deposition conditions may be to the conditions of primordial ices in planets, comets, and planetesimals in solar system formation models. The ammonia can catalyze the formation of nonstoichiometric gas hydrates which are then trapped in icy grains and accreted into planetary objects.
In addition to subsurface processes, ammonia-bearing liquids and vapors may interact directly with surface clathrate hydrate phases on Titan. In situ clathrate hydrate formations (not involving ammonia) have been considered important in determining the surface composition and structures of this moon (40–42) and the presence of ammonia may play a role in these processes.
Polar, water miscible molecules such as ammonia are traditionally considered as clathrate hydrate formation inhibitors. Our experiments and computations show that the inhibition effect is not due to inherent instability of the solid clathrate hydrate phase of these guest molecules, but is rather the stabilizing effect of these guests on the aqueous phase from which the clathrate hydrate is formed.
Materials and Methods
Sample Preparation.
Vapor codepositions of water and guest materials (ammonia, methane, and 50 mole % ammonia/methane mixture) were performed in an evacuated chamber at approximately 0.01 mbar. The water was vaporized from a pipette into the evacuation chamber by low pressure and deposited on a copper plate which was cooled down to approximately 16 K by DISPLEX (Air Products). In each experiment a total of 600 mmHg of the guest gas (ammonia, methane or 1∶1 ammonia∶methane mixture) was injected into a 1 L bulb and slowly injected into the deposition chamber. The water and guest gas vapor flows were controlled by leak valve and after 24 h approximately 2 g of water and 0.5 g of guest vapors were deposited on the copper plate. The copper plate covered with the product deposit was removed from the evacuation chamber and quickly immersed in liquid nitrogen with minimal exposure to air. Significant moisture condensation on the sample during the short transfer period is not expected. For the various characterization experiments the materials were transferred to the powder X-ray diffraction and Raman instruments.
In the single crystal experiments, a THF + NH3 binary sII hydrate was prepared from a 1∶17 solution of THF in 5% aqueous ammonia. The synthesis conditions give an approximate mole ratio of 1∶0.946∶17 for THF∶NH3∶H2O in this solution. Upon cooling the solution to 263 K masses of crystals appeared. Crystals suitable for diffraction were identified by examination under a microscope mounted in a cold box. The structure obtained from the single crystal diffraction experiment was used to determine that 39% of the small cages were occupied by ammonia which gives an approximate clathrate hydrate composition of 1∶0.78∶17 for THF∶NH3∶H2O. Different ammonia small cage occupancies are likely to be found starting from different initial compositions. Whether small cages could be filled with a more ammonia-rich liquid is uncertain and should be the subject of a separate study.
X-Ray Data Collection and Structure Solution.
Single crystal X-ray diffraction data were measured on a Bruker Apex 2 Kappa diffractometer at 100 K, using graphite monochromatized Mo Kα radiation (λ = 0.71073 Å). The unit cell was determined from randomly selected reflections obtained using the Bruker Apex2 automatic search, center, index, and least squares routines. Integration was carried out using the program SAINT, and an absorption correction was performed using SADABS (43). The crystal structures were solved by direct methods and the structure was refined by full-matrix least-squares routines using the SHELXTL program suite (44). All atoms were refined anisotropically. Hydrogen atoms on guest molecules were placed in calculated positions and allowed to ride on the parent atoms.
Powder X-Ray Diffraction (PXRD).
The PXRD pattern was recorded on a BRUKER AXS model D8 Advance diffractometer in the θ/θ scan mode using Cu Kα radiation (λ1 = 1.5406 Å, λ2 = 1.5444 Å, and I2/I1 = 0.5). The samples stored in liquid nitrogen were quickly transferred to the X-ray stage cooled down to approximately 100 K in air and diffraction patterns were measured with increasing temperature. The experiment was carried out in the step mode with a fixed time of 2 s and a step size of 0.03014° for 2θ = 8 to 42° at each temperature. The powder pattern was refined by the Le Bail method using the profile matching method within FULLPROF (45).
Raman Spectroscopy.
The Raman spectra were measured using a dispersive Raman microscope instrument with the focused 514.5 nm line of an Ar-ion laser for excitation. The scattered light was dispersed using a single-grating spectrometer (1,200 grating) and detected with a CCD detector (-90 °C, cooled by liquid nitrogen). The sample kept in liquid nitrogen was loaded onto a brass sample holder, which was cooled down to approximately 100 K in air and measured with increasing temperature. The sample holder was covered by quartz glass and dry nitrogen gas was blown over the surface during measurements in order to keep ice from condensing.
Computational Methods.
Molecular Dynamics simulations were performed with DL_POLY version 2.20 (46) using the leapfrog algorithm with a time step of 1 fs. The positions of the water oxygen atoms of the structure I and structure II clathrate hydrate phases were taken from X-ray crystallography and the water hydrogen atoms assigned to oxygen atoms in the unit cell in such a way as to simultaneously satisfy the ice rules and minimize the net unit cell dipole moment. The centers of mass of the guests are initially put in the center of the hydrate cages. For simulations of the sI clathrate hydrate, 3 × 3 × 3 replicas of the unit cell are used with 162 large cages and 54 small cages in the simulation cell. For simulations of the sII clathrate hydrate, a 2 × 2 × 2 replica of the unit cell is used with 64 large cages and 128 small cages.
In the simulations, the van der Waals and electrostatic intermolecular interactions of water are modeled with the TIP4P potential (47), THF with the AMBER force field (48) NH3 with the OPLS force field of Rizzo and Jorgensen (49), and the TRAPP potential for CH4 (50). The potential parameters in the force fields are given in Table S1. The partial electrostatic charges on the atoms of THF are calculated using the CHELPG method (51) and the other guest molecules include partial atomic charges as part of the force field. The water and guest molecules are given freedom of motion in the simulations but their internal structure is considered rigid. A cutoff of 13 Å is used for the long-range forces in the simulation.
For the THF (in large cages) + NH3 (in small cages) binary sII clathrate hydrate three sets of simulations at 100 K, (temperature of experimental single crystal structure determination), 183 K, and 200 K were performed. For each simulation, one set of simulations with the all small cages occupied by NH3 guests and a second set with 39% of the small cages occupied by NH3 guests (in accordance to the experimental values) were performed. In addition, seven simulation sets were performed at 100, 120, 140, 160, 170, 200, and 240 K for the pure NH3 sI and sII clathrate hydrates and the CH4 (large cages) + NH3 (small cages) binary sII clathrate hydrate to study the stability of these phases. Simulations for each clathrate hydrate are performed for 6 ns with 100 ps of temperature scaled equilibration to bring the clathrate hydrate with each guest combination to the desired volume and temperature in the constant pressure—constant temperature (NPT) ensemble.
The hydrogen bonding probabilities between NH3 guest and host water, p, were calculated from the radial distribution functions (RDF), g(r) between the corresponding atoms on the guest ammonia and cage water molecules,
![]() |
[1] |
In Eq. 1, rmin is the first minimum in the RDF which appears after the hydrogen bond peak of ∼2 Å which represent hydrogen bonded NH3 with the cage water molecules.
Supplementary Material
ACKNOWLEDGMENTS.
We thank G. L. McLaurin and S. Lang for technical support. We thank an anonymous reviewer for providing constructive comments on this manuscript. We also thank the National Research Council of Canada for financial support.
Footnotes
The authors declare no conflict of interest.
This article is a PNAS Direct Submission.
This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1205820109/-/DCSupplemental.
Data deposition: The single-crystal XRD structure of the binary tetrahydrofuran and ammonia structure II clathrate hydrate has been deposited in the Cambridge Structural Database, Cambridge Crystallographic Data Centre, Cambridge CB2 1EZ, United Kingdom (CSD reference no. 864781).
References
- 1.Lunine JI, Stevenson DJ. Clathrate and ammonia hydrates at high pressure: Application to the presence of methane on Titan. Icarus. 1987;70:61–77. [Google Scholar]
- 2.Fortes AD, Choukroun M. Phase behavior of ice and hydrates. Space Sci Rev. 2010;153:185–218. [Google Scholar]
- 3.Loveday JS, et al. Stable methane hydrate above 2 GPa and the source of Titan’s atmospheric methane. Nature. 2001;410:661–663. doi: 10.1038/35070513. [DOI] [PubMed] [Google Scholar]
- 4.Tobie G, Lunine JI, Sotin C. Episodic outgassing as the origin of atmospheric methane on Titan. Nature. 2006;440:61–64. doi: 10.1038/nature04497. [DOI] [PubMed] [Google Scholar]
- 5.Choukroun M, Grasset O, Tobie G, Sotin C. Stability of methane clathrate hydrates under pressure: Influence on outgassing processes of methane on Titan. Icarus. 2010;205:581–593. [Google Scholar]
- 6.Cable ML, et al. Titan tholins: Simulating titan organic chemistry in the Cassini-Huygens era. Chem Rev. 2012;112:1882–1909. doi: 10.1021/cr200221x. [DOI] [PubMed] [Google Scholar]
- 7.Fortes AD. Titan’s internal structure and the evolutionary consequences. Planet Space Sci. 2012;60:10–17. [Google Scholar]
- 8.Castillo-Rogez J, Lunine JI. Evolution of Titan’s rocky core constrained by Cassini observations. Geophys Res Lett. 2010;37:L20205. [Google Scholar]
- 9.Choukroun M, Tobie G, Grasset O. Ammonia, a methane hydrate inhibitor—Implications for Titan’s atmospheric methane; Lunar and Planetary Science, XXXVII; Papers Presented to the Thirty-Seventh Lunar and Planetary Conference; Houston, TX: 2006. p. 1640. [Google Scholar]
- 10.Sloan ED, Koh CA. Clathrate Hydrates of Natural Gases. 3rd Ed. Boca Raton, FL: CRC Press; 2008. pp. 230–232. [Google Scholar]
- 11.Uras N, Devlin JP. Rate study of ice particle conversion to ammonia hemihydrate: Hydrate crust nucleation and NH3 diffusion. J Phys Chem A. 2000;104:5770–5777. [Google Scholar]
- 12.Uras N, Buch V, Devlin JP. Hydrogen bond surface chemistry: Interaction of NH3 with an ice particle. J Phys Chem B. 2000;104:9203–9209. [Google Scholar]
- 13.Kurnosov A, Dubrovinsky L, Kuznetsov A, Dmitriev V. High Pressure/high temperature behavior of the methane—ammonia water system up to 3 GPA. Z Naturforschg. 2006;61b:1773–1576. [Google Scholar]
- 14.Nelmes RJ, Loveday JS, Guthrie M, Francis DJ. ISIS Experimental Report RB 10795. Oxfordshire, UK: Rutherford Appleton Laboratory; 2000. Molecular systems and mixtures under pressure. [Google Scholar]
- 15.Dong T, et al. Experimental determination of methane hydrate formation in presence of ammonia; Proceedings of the 6th International Conference on Gas Hydrates (ICGH 2008); Vancouver, British Columbia, Canada: 2008. Jul 6–10, [Google Scholar]
- 16.Dong T, et al. Experimental study of separation of ammonia synthesis vent gas by hydrate formation. Petrol Sci. 2009;6:188–193. [Google Scholar]
- 17.Wang L, Liu A, Guo X, Chen G, Li G. Experimental determination and calculation of methane hydrate formation in the presence of ammonia. J Chem Ind Eng (China) 2008;59:276–280. [Google Scholar]
- 18.Fortes AD, Wood IG, Brodholt JP, Vočadlo L. The structure, ordering and equation of state of ammonia dihydrate (NH3·2H2O) Icarus. 2003;162:59–73. [Google Scholar]
- 19.Fortes AD, Suard E, Lemée-Cailleau MH, Pickard CJ, Needs RJ. Crystal structure of ammonia monohydrate phase II. J Am Chem Soc. 2009;131:13508–13515. doi: 10.1021/ja9052569. [DOI] [PubMed] [Google Scholar]
- 20.Mak TCW. Hexamethylenetetramine hexahydrate: A new type of clathrate hydrate. J Chem Phys. 1965;43:2799–2805. doi: 10.1126/science.149.3680.178. [DOI] [PubMed] [Google Scholar]
- 21.Koga K, Tanaka H. Rearrangement dynamics of the hydrogen-bonded network of clathrate hydrates encaging polar guest. J Chem Phys. 1996;104:263–272. [Google Scholar]
- 22.Alavi S, Susilo R, Ripmeester JA. Linking microscopic guest properties to macroscopic observables in clathrate hydrates: Guest-host hydrogen bonding. J Chem Phys. 2009;130:174501. doi: 10.1063/1.3124187. [DOI] [PubMed] [Google Scholar]
- 23.Buch V, et al. Clathrate hydrates with hydrogen bonds. Phys Chem Chem Phys. 2009;11:10245–10265. doi: 10.1039/b911600c. [DOI] [PubMed] [Google Scholar]
- 24.Shinohara H, Nagashima U, Tanaka H, Nishi N. Magic numbers for water-ammonia binary clusters: Enhanced stability of ion clathrate structures. J Chem Phys. 1985;83:4183. [Google Scholar]
-
25.Diken EG, Hammer NI, Johnson MA, Christie RA, Jordan KD. Mid-infrared characterization of the
clusters in the neighborhood of the n = 20 “magic” number. J Chem Phys. 2005;123:164309. doi: 10.1063/1.2074487. [DOI] [PubMed] [Google Scholar] -
26.Khan A. Theoretical studies of
and NH3(H2O)20H+ clusters. Chem Phys Lett. 2001;338:201–207. [Google Scholar] -
27.Willow SY, Singh NJ, Kim KS.
Resides inside the water 20-mer cage as opposed to H3O+, which resides on the surface: A first principles molecular dynamics simulation study. J Chem Theory Comput. 2011;7:3461–3465. doi: 10.1021/ct200486c. [DOI] [PubMed] [Google Scholar] - 28.Jeffrey GA. In: Hydrate Inclusion Compounds in Inclusion Compounds. Atwood JL, Davies JED, MacNicol DD, editors. Vol 1. London: Academic Press; 1984. [Google Scholar]
- 29.Hallbrucker A, Mayer E. Unexpectedly stable nitrogen, oxygen, carbon monoxide and argon clathrate hydrates from vapor-deposited amorphous solid water: An X-ray and two-step differential scanning calorimetry study. J Chem Soc Faraday Trans. 1990;86:3785–3792. [Google Scholar]
- 30.Ripmeester JA, Ding L, Klug DD. A clathrate hydrate of formaldehyde. J Phys Chem. 1996;100:13330–13332. [Google Scholar]
- 31.Monreal IA, Devlin JP, Maşlakci Z, Çiçek MB, Uras-Aytemiz N. Controlling nonclassical content of clathrate hydrates through the choice of molecular guests and temperature. J Phys Chem A. 2011;115:5822–5832. doi: 10.1021/jp109620b. [DOI] [PubMed] [Google Scholar]
- 32.Schicks JM, Ripmeester JA. The coexistence of two different methane hydrate phases under moderate pressure and temperature conditions: Kinetic versus thermodynamic products. Angew Chem Inter Ed. 2004;43:3310–3313. doi: 10.1002/anie.200453898. [DOI] [PubMed] [Google Scholar]
- 33.Pironon J, Grimmer J, Teinturier S, Guillaume D, Dubessy J. Dissolved methane in water: Temperature effect on Raman quantification in fluid inclusions. J Geochem Explor. 2003;78–79:111–115. [Google Scholar]
- 34.Hernandez J, Uras N, Devlin JP. Coated ice nanocrystals from water-adsorbate vapor mixtures: Formation of ether—CO2 clathrate hydrate nanocrystals at 120 K. J Phys Chem B. 1998;102:4526–4535. [Google Scholar]
- 35.Delzeit L, Powell K, Uras N, Devlin JP. Ice surface reactions with acids and bases. J Phys Chem B. 1997;101:2327–2332. [Google Scholar]
- 36.Nelson RM, et al. Saturn’s Titan: Surface change, ammonia and implications for atmospheric and tectonic activity. Icarus. 2009;199:429–441. [Google Scholar]
- 37.Porco CC, et al. Cassini observes the active south pole of Enceladus. Science. 2006;311:1393–1401. doi: 10.1126/science.1123013. [DOI] [PubMed] [Google Scholar]
- 38.Fortes AD, Grindrod PM, Trickett SK, Vočadlo L. Ammonium sulfate on Titan: Possible origin and role in cryovolcanism. Icarus. 2007;188:139–153. [Google Scholar]
- 39.Fortes AD. Metasomatic clathrate xenoliths as a possible source for the south polar plumes of Enceladus. Icarus. 2007;191:743–748. [Google Scholar]
- 40.Fortes AD, Stofan ER. Clathrate formation in the near-surface environment of Titan; 36th Lunar and Planetary Science Conference; League City, TX: 2005. Mar 14–18th, (abstr 1123) [Google Scholar]
- 41.Osegovic JP, Max MD. Compound clathrate hydrate on Titan’s surface. J Geophys Res. 2005;110:E08004. [Google Scholar]
- 42.Choukroun M, Sotin C. Is Titan’s shape caused by its meteorology and carbon cycle? Geophys Res Lett. 2012;39:L04201. [Google Scholar]
- 43.Sheldrick GM. SADABS Version 2.03. Germany: University of Gottingen; 2002. [Google Scholar]
- 44.Sheldrick GM. SHELXTL, Version 6.10. Madison, WI: Bruker AXS Inc.; 2000. [Google Scholar]
- 45.Rodriguez-Carvajal J. Abstracts of the Satellite Meeting on Powder Diffraction of the XV Congress of the IUCr. Toulouse, France: International Union of Crystallography; 1990. FULLPROF: A program for rietveld refinement and pattern matching analysis; p. 127. [Google Scholar]
- 46.Smith W, Forester TR, Todorov IT. DL_POLY 2.20. Daresbury: CCLRC, Daresbury Laboratory; [Google Scholar]
- 47.Abascal JLF, Sanz E, Garcia Fernandez R, Vega C. A potential model for the study of ices and amorphous water: TIP4P/Ice. J Chem Phys. 2005;122:234511. doi: 10.1063/1.1931662. [DOI] [PubMed] [Google Scholar]
- 48.Cornell WD, et al. A second generation force field for simulation of proteins, nucleic acids, and organic molecules. J Am Chem Soc. 1995;117:5179–5197. [Google Scholar]
- 49.Rizzo RC, Jorgensen WL. OLPS all-atom model for amines: Resolution of the amine hydration problem. J Am Chem Soc. 1999;121:4827–4836. [Google Scholar]
- 50.Martin MG, Siepmann JI. Transferable potentials for phase equilibria. 1. United atom description of n-alkanes. J Phys Chem B. 1998;102:2569–2577. [Google Scholar]
- 51.Breneman CM, Wiberg KB. Determining atom-centered monopoles from molecular electrostatic potentials. The need for high sampling density in formamide conformational analysis. J Comput Chem. 1990;11:361–373. [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.





