Anda di halaman 1dari 4

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/283475268

Kinetics of methane hydrate dissociation

Article in Doklady Physical Chemistry · October 2015


DOI: 10.1134/S0012501615100061

CITATION READS

1 120

2 authors, including:

S. Ya. Misyura
Russian Academy of Sciences
59 PUBLICATIONS 391 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

1. Influence of of structures on crystallization kinetics View project

Petroleum Waxes Deposition and Removal View project

All content following this page was uploaded by S. Ya. Misyura on 05 November 2015.

The user has requested enhancement of the downloaded file.


ISSN 00125016, Doklady Physical Chemistry, 2015, Vol. 464, Part 2, pp. 244–246. © Pleiades Publishing, Ltd., 2015.
Original Russian Text © V.E. Nakoryakov, S.Ya. Misyura, 2015, published in Doklady Akademii Nauk, 2015, Vol. 464, No. 6, pp. 693–695.

PHYSICAL
CHEMISTRY

Kinetics of Methane Hydrate Dissociation


Academician V. E. Nakoryakov and S. Ya. Misyura
Received July 2, 2015

Abstract—The methane hydrate dissociation at an external pressure of 1 bar has been experimentally studied.
The methane dissociation rate depends on the heat transfer intensity in the course of dissociation. Quasiiso
thermal and nonisothermal conditions of clathrate dissociation have been compared. Under nonisothermal
conditions, the dissociation rate depends on the heat flux density. The lowest heat flux was achieved in a
quasistationary mode, and the highest heat flux was achieved through combustion of methane released from
the clathrate powder. The 54fold increase in heat flux leads to the ninefold increase in the methane dissoci
ation rate. Depending on the heat flux density, the following variants of clathrate dissociation were observed:
without selfpreservation, partial selfpreservation with one dissociation rate minimum, and partial selfpres
ervation with two minima.
DOI: 10.1134/S0012501615100061

Formation and dissociation of gas hydrate [1] ously increased from the initial liquid nitrogen tem
depend on a large number of factors: pressure, tem perature to the melting point of ice at ambient pres
perature, subcooling degree, and structural character sure. Methane hydrate had an sI structure.
istics of a sample. The dissociation kinetics has been Characteristics of the methane hydrate powder were
studied in [2] where the clathrate decomposition rate determined by gravimetric and volumetric methods.
as a function of gas volatility has been determined. The Methane hydrate decomposition leads to the
kinetics of dissociation at various heating rates has appearance of a multilayer structure (Fig. 1a). With a
been studied in [3]. To date, available experimental lapse of time, the ice crust thickness δ increases, and
data mainly pertain to clathrate decomposition in low the ice crust radius Ri decreases. Measurements with a
heat fluxes. In studies dealing with methane combus scanning electron microscope [11] have shown that
tion [4–7], heat fluxes have not been measured and clathrate decomposition leads to the formation of
local rates of methane hydrate decomposition have not pores in ice. Gas diffusion through the pores exceeds
been determined. Therefore, quantitative data on the by many orders of magnitude the diffusion through the
decomposition kinetics in combustion are of interest. solid crust (Fig. 1b). In the course of dissociation, a gas
Special attention focuses on the phenomenon of “self hydrate can exhibit anomalously slow decomposition
preservation” [8, 9] where extremely low clathrate at annealing temperatures, i.e., selfpreservation. In
decomposition rates are observed. Currently, we are this case, the pores are closed with ice, and the crust
still far from a deep understanding of the selfpreserva becomes continuous, as shown in Fig. 1a. Crank [12]
tion mechanism. Anomalous stability is attributed to has derived the equation that relates the change in ice
the formation of the solid ice crust and to the appear layer thickness (without pores) in time with diffusion
ance of characteristic texture in the course of anneal and pressure.
ing. Quantitative data on the growth of ice crust as a A = 3(1 – R2) + 2(R3 – 1) = c1D(Pd(T) – Pa)τ, (1)
function of time have been reported [10]. The present
paper deals with methane hydrate dissociation mech where с1 is a constant; Pd(T) is the methane dissocia
anisms and selfpreservation. tion pressure at temperature T; Pa is the ambient
methane dissociation pressure; D is the methane diffu
The scheme of the experimental setup was sion coefficient through the surface ice layer; R =
described in [6]. The initial weight concentration of Ri/R0, where R0 is the initial radius of the forming ice
methane in a gas hydrate was 12.1 wt %, and the initial crust, which corresponds to the grain (sphere) radius,
average diameter of powder grains was d = 2.1–2.6 mm. and Ri is the current ice crust radius; τ is time. Analysis
The gas hydrate dissociation occurred under noniso of experimental data has shown that the ice layer
thermal conditions: the sample temperature continu thickness δ actually becomes equal to the sphere
radius R0 at the moment of time τ = τ1 = (0.3–0.5)τmax
(Fig. 1c), where τ1 corresponds to the complete tran
Kutateladze Institute of Thermophysics, Siberian Branch, sition of the gas hydrate to ice. The powder surface
Russian Academy of Sciences, pr. Akademika Lavrent’eva 1, temperature for a given segment is 270–273 K, and the
Novosibirsk, 630090 Russia temperature curve is linear. The grain surface temper
email: misura@itp.nsc.ru ature was measured by means of a heat imager. Such a

244
KINETICS OF METHANE HYDRATE DISSOCIATION 245

(a) (b) T, °C (c)

2 0

Ri 1
τ1 τmax

Fig. 1. (a) A multilayer structure formed during the gas hydrate dissociation: 1 is the outer ice shell, 2 is the gas layer under ice,
and 3 is the inner gas hydrate core. (b) Methane release from pores formed in the ice shell. (c) Temperature of the grain surface
as a function of time.

long time interval from τ1 to τmax is caused by a low realized for curve 1. Usually, upon the phase transi
temperature gradient and insignificant heat flux rather tion, the powder temperature decreases by several
than by the phase transition. All gas hydrate converts kelvins. At higher heat fluxes, only one dissociation
into ice long before all methane is released, which is minimum is observed. It should be emphasized that,
caused by small pore diameters (on the order of 0.1– when the second minimum is observed under noniso
1.0 µm) and the high resistance to gas passage through thermal conditions, the ice crust is rather thick, and 10
the pores. The crystallization front velocity exceeds to 40% of methane is retained in the sample [14].
the average gas outflow rate through the pores. Gas Under quasiisothermal conditions (curve 1), decom
outflow is accompanied by a decrease in the gas pres
position begins just at annealing temperature (the
sure in the pores, so that the gas flow rate decreases.
Experimental methane hydrate dissociation curves external pressure rapidly decreases to 1 bar), and self
when a solid ice crust is formed, i.e., under selfpres preservation occurs for a very thin ice film. Before the
ervation conditions, have been obtained [10]. It has onset of selfpreservation, the sample can lose only a
been established that even at very low decomposition few percent of the methane weight.
rates (ΔV/V) on the order of 0.1% per hour, the diffu Thus, under nonisothermal conditions, the meth
sion coefficients are many orders of magnitude higher ane hydrate dissociation rate depends on the heat flux
than the vacancy diffusion coefficients. Evaluation
density. The 54fold increase in the heat flux density
also indicates a substantial decrease in activation
energy. Thus, it is more correct to speak about partial led to the ninefold increase in the dissociation rate.
(rather than complete) selfpreservation in which even Depending on the heat flux density, different variants
a very low pore density plays a decisive role in the of gas hydrate decomposition can be observed with
mechanism of gas hydrate dissociation. The diffusion
coefficient in Eq. (1) shall be a function of defect con
tent even at an extremely low pore density, and Eq. (1) ν, %/s
shall be nonlinear (D = var). Figure 2 shows the meth 1
ane hydrate dissociation curves at the atmospheric 0.1
pressure 1 bar. The high heat flux (curve 3) was
achieved through combustion of methane released 0.01
from the powder. The heat flux was determined from
the gas temperature profile near the powder surface 0.001
and was linear (low Rayleigh numbers). The low heat 0.0001
flux (curve 4) was achieved by regulating the ambient 1
1E05 2
temperature, which was 10–13°C higher than the 3
powder temperature during the entire decomposition 1E06 4
period. Curve 2 was obtained at constant ambient tem
perature 23°C, while the sample temperature was 1E07
changed from liquid nitrogen temperature to 0°C. The 180 200 220 240 260 280 300
54fold increase in the heat flux density (from 255 to T, K
13 700 W/m2) led to the ninefold increase in the max
Fig. 2. Methane hydrate dissociation: (1) quasiisothermal
imum dissociation rate. The second minimum in the conditions [13], (2) nonisothermal conditions at the average
decomposition rate curve begins to appear at q = heat flux density q = 255 W/m2, (3) for q = 13 700 W/m2
65 W/m2 or at lower values. The lowest heat flux is (methane combustion), and (4) for q = 65 W/m2.

DOKLADY PHYSICAL CHEMISTRY Vol. 464 Part 2 2015


246 NAKORYAKOV, MISYURA

one or two minima of anomalously low decomposition 6. Misyura, S.Y., Nakoryakov, V.E., and Elistratov, S.L.,
rate. J. Eng. Thermophys., 2013, vol. 22, pp. 87–92.
7. Misyura, S.Y., Nakoryakov, V.E., and Elistratov, S.L.,
J. Eng. Thermophys., 2013, vol. 22, pp. 169–173.
ACKNOWLEDGMENTS 8. Melnikov, V.P., Nesterov, A.N., Reshetnikov, A.M., and
This work was supported by the Russian Science Istomin, V.A., Chem. Eng. Sci., 2011, vol. 66, pp. 73–
Foundation (project no. 15–19–10025). 77.
9. Nakoryakov, V.E. and Misyura, S.Y., Chem. Eng. Sci.,
2013, vol. 104, pp. 1–9.
REFERENCES 10. Shimada, W., Takeya, S., Kamata, Y., Uchida, T.,
1. Istomin, V.A. and Yakushev, V.S., Gas Hydrates in Nagao, J., Ebinuma, T., and Narita, H., J. Phys. Chem. B,
Nature, Moscow: Nedra, 1992. 2005, vol. 109, pp. 5802–5807.
2. Jamaluddin, A.K.M., Kalogerakis, N., and Bishnoi, P.R., 11. Falenty, A. and Kuhs, W.F., J. Phys. Chem. B, 2009,
Can. J. Chem. Eng., 1989, vol. 67, pp. 948–954. vol. 113, pp. 15975–15988.
12. Crank, J., The Mathematics of Diffusion, 2nd ed.,
3. Sato, H., Sakamoto, H., Ogino, S., Mimachi, H.,
Oxford: Oxford Univ. Press, 1975, pp. 89–103.
Kinoshita, T., Iwasaki, T., Sano, K., and Ohgaki, K.,
Chem. Eng. Sci., 2013, vol. 91, pp. 86–89. 13. Kuhs, W.F., Genov, G., Staykova, D.K., and Hansen, T.,
Phys. Chem. Chem. Phys., 2004, vol. 6, pp. 4917–4920.
4. Takeuchi, M., Ueda, T., Amari, T., and Mizomoto, M.,
Trans. Jpn Soc. Mech. Eng., Ser. B, 1998, vol. 64, 14. Takeya, S. and Ripmeester, J.A., Angew. Chem., Int. Ed.
pp. 3485–3490. Engl., 2008, vol. 47, pp. 1276–1279.
5. Misyura, S.Y. and Nakoryakov, V.E., Energ. Fuels,
2013, vol. 27, pp. 7089–7097. Translated by G. Kirakosyan

DOKLADY PHYSICAL CHEMISTRY Vol. 464 Part 2 2015

View publication stats

Anda mungkin juga menyukai