You are on page 1of 4

Chemical Engineering Science 61 (2006) 7827 7830

www.elsevier.com/locate/ces

Shorter Communication

Notes on the dissolution rate of gas hydrates in undersaturated water


Grard C. Nihous , Stephen M. Masutani
Hawaii Natural Energy Institute, University of Hawaii, Honolulu HI 96822, USA
Received 30 June 2005; received in revised form 16 September 2005; accepted 20 September 2005

Abstract
An elementary model for the dissolution of pure hydrate in undersaturated water is proposed that combines intrinsic decomposition within a
desorption lm and the subsequent diffusion of the released hydrate guest species into bulk water. Applying the proposed approach to recently
published measurements of the decomposition rates of methane (CH4 ) and carbon dioxide (CO2 ) hydrates in deep seawater suggests that the
concentration of the hydrate guest species at the interface between desorption lm and diffusive boundary layer may be much lower than
ambient solubility. Calculations, however, fail to account for the observed proportionality of decomposition rate with solubility for both CH4
and CO2 hydrates. This may indicate a limitation in the range of applicability of published formulas for intrinsic hydrate decomposition rates.
2005 Elsevier Ltd. All rights reserved.
Keywords: Gas hydrate; Dissolution; Decomposition; Kinetics; Phase equilibria; Phase change

1. Background
Vast deposits of methane hydrates below the seaoor of most
continental margins represent an energy resource that could be
exploited in the future (Max, 2000). In some cases, hydrates
have been found lying exposed on the seaoor, in contact with
the water column (MacDonald et al., 1994). Warmer ocean waters in a changing climate may destabilize some marine hydrate
deposits and potentially generate an adverse climatic feedback
in the process (Harvey and Huang, 1995). Furthermore, sequestration of anthropogenic CO2 in the deep ocean has been proposed at conditions that foster hydrate formation. The stability
of the CO2 hydrate phase will affect the term of sequestration and extent of local environmental impact (Fer and Haugan,
2003). It is important, therefore, to understand well the phenomena governing the evolution of hydrates immersed in water
greatly undersaturated with the hydrate-forming species.
While the kinetics of hydrate dissociation have been investigated experimentally in the laboratory, the liquid phase in
contact with the hydrate was typically saturated with the guest
species, or, in the case of CO2 , the hydrate formed as a thin

Corresponding author. Tel.: +1 808 956 2338; fax: +1 808 956 2336.

E-mail address: nihous@hawaii.edu (G.C. Nihous).


0009-2509/$ - see front matter 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2005.09.010

unstable layer on liquid CO2 droplets (Teng et al., 1995; Hirai


et al., 1997). In situ measurements of the dissolution rates of
pure CH4 and CO2 hydrates in the deep ocean have only recently been published (Rehder et al., 2004).
In this communication, we propose a simple model to assess
the kinetics of hydrate dissolution in undersaturated water. This
model is then applied to interpret the data of Rehder et al.
(2004).
2. A model of hydrate dissolution in undersaturated water
Kim et al. (1987) introduced the concept of intrinsic decomposition rate to represent the breakdown of hydrate particles
within a reaction-desorption layer at the soliduid boundary.
In a series of carefully designed experiments, Clarke and Bishnoi (2000, 2001, 2004, 2005) were able to measure intrinsic
hydrate decomposition and formation rates for different guest
molecules (methane, ethane and carbondioxide). The intrinsic
decomposition rate in mol/s is expressed as follows:


dn
E
(1)
= K0 exp
A(feq fs )
dt
RT
n reects the traditional hydrate stoichiometric formula of one
mole hydrate per mole of guest molecule; K0 is a constant
in mol/(s m2 Pa), E an activation energy (J/mol), R the gas

7828

G.C. Nihous, S.M. Masutani / Chemical Engineering Science 61 (2006) 7827 7830

Fig. 1. Equilibrium phase diagram for methane hydrate, gas (CH4 -rich solution) and seawater solution (equation of Handa, 1990, for pure water with
an offset of 1.15 K following Peltzer and Brewer, 2000).

constant, T the absolute temperature (K) and A the surface of


the decomposing particles; feq is the equilibrium fugacity (Pa)
of the guest species in the hydrate lattice and fs the fugacity
in the desorption layer. The fugacity difference in Eq. (1) is
a measure of the state of disequilibrium of the hydrate, and
quanties the driving decomposition force.
Clarke and Bishnoi (2000) illustrated the process of hydrate
dissociation in their experiments by showing the decomposing
particles and their desorption lm surrounded by a gas cloud
(their Fig. 3); accordingly, the bulk liquid phase was saturated
with the gas. More generally, saturation of the water-rich phase
is implicit when hydrate stability phase diagrams similar to
Fig. 1 are shown. If heating or pressure reduction is used to
decompose hydrates, then the driving fugacity difference corresponds to a state point delimited by either f or fp . The
curve in Fig. 1, however, represents equilibrium between three
coexisting phases (hydrate, gas, liquid), and the composition
of all phases is xed once temperature or pressure is given. If
the hydrate is immersed in a sufcient volume of water greatly
undersaturated with the hydrate guest species, then it will readily decompose even though the temperature and pressure of
all phases fall within the hydrate stability regime predicted by
Fig. 1. In such cases where only two phases (hydrate and liquid)
will exist, the system has two degrees of freedom according to
the Phase Rule, one of which is associated with a state variable
other than pressure and temperature:1 the hydrate guest species
concentration in the water-rich phase.
This point is illustrated in Fig. 2, which shows the equilibrium concentration (solubility) of a hydrate-forming solute at
a given temperature. The horizontal line originating from point
T represents three-phase equilibrium, and its projection on a
pressuretemperature plane is a point belonging to the curve in
Fig. 1. Even at pressures higher than the three-phase equilibrium value, a hydrate sample will decompose when immersed in
water if the mole fraction of dissolved CH4 in the liquid phase
1 When there is no ambiguity, salinity s is not explicitly mentioned

though it does affect equilibrium.

Fig. 2. Equilibrium concentration (solubility) of CH4 in seawater with or


without hydrates at T =3.6 C(276.75 K) and a salinity of 34.6 ppt; the dotted
line corresponds to metastable solubility if hydrates did not form.

Fig. 3. Equilibrium CH4 fugacity as a function of pressure for the same


system and conditions as in Fig. 2.

falls to the left of the solubility curve; i.e., when the reactiondesorption layer is not saturated. For example, hydrate will decompose when the liquid desorption layer corresponds to Point
B, but not to Point A. A fugacity difference between the hydrate
solid surface at equilibrium (Point C) and the desorption layer
(Point B) then arises, as shown in Fig. 3. This decomposition
mechanism does not allow a gas cloud to exist around the hydrate particle and its desorption layer. Instead, two liquid lms
separate the solid surface and bulk water: the desorption layer
and a diffusive boundary layer. As discussed above, however,
the concentration at the interface between these two lms theoretically should be lower than ambient solubility (xB < xA ).
This is a point of departure from the standard assumption that
diffusion is driven from xA .
3. Interpreting methane hydrate dissolution data
A decomposition rate per unit area of 3.7 104 mol/m2 s
was measured by Rehder et al. (2004) for methane hydrate
samples in deep seawater (ambient conditions: T = 3.6 C,

G.C. Nihous, S.M. Masutani / Chemical Engineering Science 61 (2006) 7827 7830

s = 34.6 ppt; PA = 10.48 MPa). In what follows, it is assumed


that the constants K0 =36000 mol/(s m2 Pa) and E =81 kJ/mol
from Clarke and Bishnoi (2001) can be used in Eq. (1). The
fugacity difference f = (feq fs ) driving hydrate decomposition can then be calculated. We nd f = 19.94 MPa.
The equilibrium fugacity of CH4 feq for the hydrateseawater
system (e.g., the curve above Point T in Fig. 3) was obtained
with the model of Munck and Skjold-JZrgensen (1988) based
on van der Waals and Platteeuw (1959). A temperature offset
of 1.15 K was used to account for the destabilizing effect of
salinity on the hydrate lattice. To relate CH4 fugacity f and
concentration x in seawater assumed to be an ideal solution, a
Henry constant H(T, s, P) was used. H0 = H (T , s, P0 = 1 atm)
was derived from Wiesenburg and Guinasso (1979) and the
effect of pressure included with vliq = 37.3 cm3 (Akinev and
Diamond, 2003):



vliq (P P0 )
f = H x = H0 exp
x.
(2)
RT
We then have2 fs =H (PA )xB . Calling PC the unknown pressure corresponding to Point C in Figs. 2 and 3, feq (PC ) =
H (PC )xB . Since feq (PC ) = fs + f as well, elementary substitutions yield the following implicit equation in PC :



vliq (PA PC )
feq (PC ) 1 exp
f = 0.
(3)
RT
Eq. (3) was solved for the methane hydrate data of Rehder
et al. (2004). The pressure PC was found to be as large as
153.76 MPa; this can be understood from the slow increase of
the hydrate equilibrium fugacity with pressure, as shown in
Fig. 3. Of interest is the degree of undersaturation of the
reaction-desorption layer , which may be dened as the ratio
of xB over the ambient equilibrium mole fraction (at Point A
in Figs. 2 and 3). It easily can be shown that:


feq (PC )
vliq (PA PC )
=
exp
.
(4)
feq (PA )
RT
For the data under consideration, we obtain  = 0.52.
It was veried that in order to drive the dissolution of hydrates in greatly undersaturated bulk waters, the liquid desorption layer in contact with the decomposing hydrate lattice also
must be undersaturated, according to the model proposed in
the previous section. It should be mentioned that the value of 
calculated above possibly could be closer to 1.0 since the dissolution rates per unit surface reported in Rehder et al. (2004)
likely refer to the overall geometric surface of the hydrate sample; the actual surface A of the decomposing hydrate particles
could be higher, resulting in smaller values of f and PC .
One may wonder what are order-of-magnitude thicknesses
for the desorption and diffusive boundary layers, L and Z, respectively. Structure I hydrates are cubic cells of dimension
while hydrate decomposition products have a guest
 12 A,
2 The system considered here is assumed to be in isothermal equilibrium
and its salinity to remain constant; hence, the shorthand notation H (P ) and
feq (P) is adopted.

7829

mole fraction 1/(1+nH ), where nH is the hydrate number typically close to 6. If we assume that only one hydrate layer at a
time is decomposing, we should have:
L


xB (1 + nH )

(5)

In the above calculations, xB = 5.7 104 . This would yield


L 0.3 m. In quasisteady-state, the dissolution ux expressed
in Eq. (1) is equal to the diffusive ux through the diffusive
boundary layer. The CH4 concentration for ambient seawater is
nearly zero. With xB expressed in mol/m3 (as CB ), and calling
D the molecular diffusivity of the hydrate forming species,
equal to about 109 m2 /s, it follows that:
dn
ADC B

dt
Z

(6)

In the above case, (dn/dt)/A = 3.7 104 mol/m2 s and CB =


31.7 mol/m3 ; Eq. (6) would yield Z 86 m, or a diffusive
layer two orders of magnitude thicker than L.
4. Interpreting CO2 hydrate dissolution data
A decomposition rate per unit area of 4.15 103 mol/m2 s
was also reported by Rehder et al. (2004) for CO2 hydrate
samples immersed in the same ambient deep seawater as the
methane hydrate samples. The authors noted that, when comparing the two hydrate types, the ratio of the dissolution rates
matched the ratio of the ambient solubilities of the respective
hydrate guest species:
(1/A)(dn/dt)(CH4 )
CA (CH4 )

(1/A)(dn/dt)(CO2 )
CA (CO2 )

(7)

Since the parameters D and Z are similar for CH4 and CO2 , they
explained this result by considering diffusive boundary layers
where diffusion takes place from saturated concentrations (with
near zero ambient concentrations), i.e., Eq. (6) for either species
but with a concentration equal to the solubility CA . If instead
the hydrate desorption lm is undersaturated, then the ratio of
the dissolution rates should match the ratio of the interfacial
concentrations CB by the same argument:
(1/A)(dn/dt)(CH4 )
CB (CH4 )

.
(1/A)(dn/dt)(CO2 )
CB (CO2 )

(8)

Eqs. (7) and (8) together suggest that the degree of undersaturation  is about the same for the two types of decomposing
hydrates.
Clarke and Bishnoi (2004) measured the intrinsic decomposition parameters for CO2 hydrates, K0 = 1.83
108 mol/(s m2 Pa) and E = 102.88 kJ/mol, respectively. The
analysis done in the previous section was repeated using those
values. To relate fugacity to mole fraction as in Eq. (2), the
model of Edwards et al. (1978) for the equilibrium of volatile
weak electrolytes in non-ideal solutions was used in combination with the seawater solubility data of Weiss (1974). Here,
vliq = 32.8 cm3 (Akinev and Diamond, 2003). We found a
very high equilibrium pressure PC and a desorption layer

7830

G.C. Nihous, S.M. Masutani / Chemical Engineering Science 61 (2006) 7827 7830

with a much greater degree of unsaturation (  0.52). This


contradicts deep-sea observations.
We postulated instead, for the sake of argument, that  for
the dissolving CO2 hydrates is about 0.52. We could then perform the same calculations as above backwards, without specifying, for example, the activation parameter E in Eq. (1). We
obtained xs =7.87103 , feq (PA )=1.96 MPa, fs =1.02 MPa,
feq (PC )=37.90 MPa and f =36.88 MPa. With the same value
of K0 given by Clarke and Bishnoi (2004), Eq. (1) yielded a
hypothetical value E=96.49 kJ/mol. This is 6.39 kJ/mol (about
6%) lower than the value measured by Clarke and Bishnoi
(2004). It shows the sensitivity of results on E, in the exponential Arrhenius-like term in Eq. (1) and illustrates a possible issue
with the applicability of Eq. (1). Parameters such as K0 and
E were measured in hydrate decomposition experiments based
on a modest depressurization. It is possible that the values of
the parameters or even the formalism embodied in Eq. (1) may
not be adequate in some instances of hydrate decomposition.
5. Conclusions
The formalism describing the dissociation kinetics of
hydrates as an intrinsic reaction in a thin liquid desorption
layer was extended to cases when hydrates are immersed in
undersaturated water. The simple model rests on the fact that
two-phase equilibrium then governs hydrate dissociation, with
the concentration of the hydrate forming species in water playing a key role in providing the necessary fugacity gradient,
beside temperature and pressure: in other words, and unlike
instances of three-phase equilibrium, neither heating nor depressurization is necessary to trigger hydrate decomposition.
An application of the proposed concept to the methane
hydrate data of Rehder et al. (2004) established that the
desorption layer itself should be undersaturated for hydrate
dissolution to proceed (when pressure and temperature do not
vary). When dealing with CO2 data, however, an application
of the model using published kinetic parameters for hydrate
dissociation failed to explain some of the eld observations.
This may indicate that published formulas for intrinsic
hydrate decomposition rates are not applicable in the conditions of the eld experiment.
The ideas developed in this paper clearly should be further
tested by ad hoc experiments.
Acknowledgements
This work was supported by a Grant from the Ofce of Naval
Research and was conducted as part of the Hawaii Energy and
Environmental Technology initiative. Also, the authors would
like to thank two anonymous reviewers for their critical insights.

References
Akinev, N.N., Diamond, L.W., 2003. Thermodynamic description of aqueous
nonelectrolytes at innite dilution over a wide range of state parameters.
Geochimica et Cosmologica Acta 67 (4), 613627.
Clarke, M., Bishnoi, P.R., 2000. Determination of the intrinsic rate of ethane
gas hydrate decomposition. Chemical Engineering Science 55, 48694883.
Clarke, M., Bishnoi, P.R., 2001. Determination of the activation energy and
intrinsic rate constant of methane gas hydrate decomposition. Canadian
Journal of Chemical Engineering 79, 143147.
Clarke, M., Bishnoi, P.R., 2004. Determination of the intrinsic rate constant
and activation energy of CO2 gas hydrate decomposition using in-situ
particle size analysis. Chemical Engineering Science 59, 29832993.
Clarke, M., Bishnoi, P.R., 2005. Determination of the intrinsic kinetics of
CO2 gas hydrate formation using in-situ particle size analysis. Chemical
Engineering Science 60, 695709.
Edwards, T.J., Maurer, G., Newman, J., Prausnitz, J.M., 1978. Vaporliquid
equilibria in multicomponent aqueous solutions of volatile weak
electrolytes. Journal of the American Institute of Chemical Engineers 24
(6), 966976.
Fer, I., Haugan, P.M., 2003. Dissolution from a liquid CO2 lake disposed in
the deep ocean. Limnology and Oceanography 48 (2), 872883.
Handa, Y.P., 1990. Effect of hydrostatic pressure and salinity on the stability
of gas hydrates. Journal of Physical Chemistry 94, 26522657.
Harvey, L.D.D., Huang, Z., 1995. Evaluation of the potential impact of
methane clathrate destabilization on future global warming. Journal of
Geophysical Research 100 (D2), 29052926.
Hirai, S., Okazaki, K., Tabe, Y., Hijikata, K., Mori, Y., 1997. Dissolution rate
of liquid CO2 in pressurized water ows and the effect of clathrate lms.
Energy 22, 285293.
Kim, H.C., Bishnoi, P.R., Heidemann, R.A., Rizvi, S.S.H., 1987. Kinetics
of methane hydrate decomposition. Chemical Engineering Science 42,
16451653.
MacDonald, I.R., Guinasso, N.L., Sassen Jr., R., Brooks, J.M., Lee, L., Scott,
K.T., 1994. Gas hydrate that breaches the sea oor on the continental
slope of the Gulf of Mexico. Geology 22, 699702.
Max, M.D. (Ed.), 2000. Natural Gas Hydrate in Oceanic and Permafrost
Environments, Kluwer Academic Publishers, Dordrecht, 415pp.
Munck, J., Skjold-JZrgensen, S., 1988. Computations of the formation of gas
hydrates. Chemical Engineering Science 43 (10), 26612672.
Peltzer, E.T., Brewer, P.G., 2000. Practical physical chemistry and empirical
predictions of methane hydrate stability. In: Max, M.D. (Ed.), Natural
Gas Hydrate in Oceanic and Permafrost Environments. Kluwer Academic
Publishers, Dordrecht, pp. 1728.
Rehder, G., Kirby, S.H., Durham, W.B., Stern, L.A., Peltzer, E.T., Pinkston, J.,
Brewer, P.G., 2004. Dissolution rates of pure methane hydrate and carbondioxide hydrate in undersaturated seawater at 1000-m depth. Geochimica
et Cosmologica Acta 68 (2), 285292.
Teng, H., Kinoshita, C.M., Masutani, S.M., 1995. Hydrate formation on the
surface of a CO2 droplet in high-pressure/low-temperature water. Chemical
Engineering Science 50 (4), 559564.
van der Waals, J.H., Platteeuw, J.C., 1959. Clathrate solutions. In: Prigogine,
I. (Ed.), Advances in Chemical Physics. Interscience, pp. 257.
Weiss, R.F., 1974. Carbon dioxide in water and seawater: the solubility of a
non-ideal gas. Marine Chemistry 2, 203215.
Wiesenburg, D.A., Guinasso Jr., N.L., 1979. Equilibrium solubilities of
methane carbon monoxide and hydrogen in water and sea water. Journal
of Chemical and Engineering Data 24 (4), 356360.

You might also like