You are on page 1of 8

International Journal of Heat and Mass Transfer 92 (2016) 766773

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


journal homepage: www.elsevier.com/locate/ijhmt

Analysis of heat transfer influences on gas production from methane


hydrates using a combined method
Yongchen Song, Jiaqi Wang, Yu Liu, Jiafei Zhao
School of Energy and Power Engineering, Dalian University of Technology, Dalian 116024, PR China

a r t i c l e

i n f o

Article history:
Received 27 May 2015
Received in revised form 13 July 2015
Accepted 17 August 2015
Available online 30 September 2015
Keywords:
Methane hydrate
Combined method
Conduction
Convection
Injection temperature
Heat transfer

a b s t r a c t
Heat transfer affects the pressure and temperature distributions of hydrate sediments, thereby controlling hydrate dissociation. Therefore, its study is essential for planning hydrate exploitation. Previously,
a two-dimensional axisymmetric model, to investigate the influence of heat transfer on hydrate exploitation from hydrate-bearing sediments, was developed and verified. Here, we extended our investigation to
the influence of heat transfer on methane gas production using a combined method coupling depressurization and thermal stimulation. Our simulations showed that during decomposition by the combined
method, a high specific heat capacity of the hydrate-bearing porous media or a high initial water content
could inhibit gas generation. However, the initial water content had only a weak influence on the cumulative gas production and generation rate. The influence of water and methane heat convection was also
weak. An increase of the thermal conductivity initially inhibited hydrate dissociation but later promoted
it. The implementation of the combined method increased gas generation compared with using only thermal stimulation. However, the benefits gradually diminished with an increasing heat injection
temperature.
2015 Elsevier Ltd. All rights reserved.

1. Introduction
Worldwide, a large volume of hydrates can be found beneath
the sea and in permafrost sediments [1]. Considering their potential as energy resources, the development of safe and efficient
methods for gas extraction from gas hydrate sediments has
become a widespread aim. At present, different methods for
exploitation have been proposed, typically based on the disruption
of thermodynamic equilibrium by depressurization [25], thermal
stimulation [68], or a combined method [9,10].
For a safe and effective exploitation of hydrates and to avoid the
limitations and disadvantages of a single approach, methods combining multiple techniques have recently been developed. Liu et al.
[11] developed a one-dimensional mathematical model to predict
hydrate decomposition in hydrate sediments via depressurization
and thermal stimulation. Their simulations showed that thermal
stimulation at constant temperature plays a limited role in hydrate
exploitation compared with depressurization. Li et al. [12] conducted an experimental study to investigate whether the combination of thermal stimulation and depressurization was propitious to
natural gas hydrate dissociation and their results suggested that
such combination could achieve a higher energy efficiency. Bai
Corresponding author. Tel./fax: +86 411 84706722.
E-mail address: jfzhao@dlut.edu.cn (J. Zhao).
http://dx.doi.org/10.1016/j.ijheatmasstransfer.2015.08.102
0017-9310/ 2015 Elsevier Ltd. All rights reserved.

and Li [13] used physical and mathematical models based on this


combined method to analyze how gas and water production were
influenced by multiphase fluid flow, kinetic and endothermic processes during decomposition, and heat convection and conduction.
Their simulations showed that, under certain conditions, the combined method provided a longer and more stable period of high gas
extraction rates over the single method. Feng et al. [1416] used a
one-dimensional system for depressurization and thermal stimulation experimental studies. To date, studies typically indicated that
the combined method is more advantageous for hydrate exploitation than a single production method. However, few studies analyzed the effect of heat transfer.
Heat transfer affects the pressure and temperature distributions
in hydrate sediments, thereby controlling hydrate decomposition
[1]. This study extends our previous investigation [17,18] to
address the influence of heat transfer on hydrate exploitation using
the combined method. We focused on the various heat transfer
modes that affect the gas generation rate and cumulative production, including the sensible heat, conductive heat flow, and convective heat transfer.
2. Modeling methodology
The mathematical models and assumptions made in this study
were based on our previous work [1721]. Three components (gas,

Y. Song et al. / International Journal of Heat and Mass Transfer 92 (2016) 766773

767

Fig. 1. Scheme of the computational hydrate core sample adapted from previous work [1721].

Fig. 2. Cumulative gas production over time for different stimulation temperatures. The thermal stimulation and combined methods were compared for different
temperatures.

water, and hydrate) and three phases (gas, liquid, and solid) were
represented in the models. Equations for the mass conservation,
energy conversion, reaction kinetics, motion, and state of the three
components were used to simulate hydrate dissociation from
hydrate reservoirs. In the model, the core was immersed in a water

bath, with an outlet valve located on the left side of the core. The
walls and the right side of the core were considered no-slip boundaries (Fig. 1). Free convection heat transfer was assumed between
the circular wall and the surroundings. Adiabatic boundary conditions were imposed at the ends of the core. These conditions were

Fig. 3. (a) Gas generation rate and (b) cumulative gas production over time for different core specific heat capacities (Cps): 0, 0.8, and 1.6 kJ/(kg K).

768

Y. Song et al. / International Journal of Heat and Mass Transfer 92 (2016) 766773

Fig. 4. Simulated core temperatures after 3 min for core specific heat capacities (Cps) of (a) 0.0, (b) 0.8, and (c) 1.6 kJ/(kgK).

based on the initial conditions of Masuda et al. [22] and described


as:

T T 0 ; P P0 ; Sh Sh0 ; Sw Sw0 ; Sg Sg0 ;

0 6 r 6 R; 0 6 x 6 L

where T is the temperature, P is the pressure, Sh is the hydrate saturation, Sw is the water saturation, and Sg is the gas saturation. The
subscript 0 indicates the initial conditions. R, r, x, and L correspond
to the core radius, the variable of core radius, the variable of core
length and core length, respectively.
The following boundary conditions were imposed:

P P0 x 0; @P=@x 0 x L; @P=@r 0 r 0; R0 ;
@T=@r 0 r 0; and @T=@x 0 x 0; L:
The specific heat capacity (Cps) of the core material was
assigned as 0, 0.8, or 1.6 kJ/(kgK), which reflects values for typical
sand and sandy clay soils ranging from 0.83 to 1.38 kJ/(kgK). The
thermal conductivity (ks) of the core material was assumed as 1.5
or 8 W/(mK), which is similar to the thermal conductivities
of vitreous silica and quartz [23], respectively. The convection
heat transfer of water (nw) and gas (ng) controls the relative
permeability of water (k rw ) and gas (krg ) expressed as
krw Sw =Sw Sg  Swr =1  Swr  Sgr ^ nw and krg Sg =Sw
Sg  Sgr =1  Swr  Sgr ^ ng , respectively, where Swr is the

relative water saturation and Sgr is the relative gas saturation. For
the basic case, low relative gas permeability, and high relative
water permeability, we set nw = 4 and ng = 2, nw = 4 and ng = 4,
and nw = 2 and ng = 2, respectively.
3. Results and discussion
Although it is not essential to heat the sediments during
hydrate decomposition by depressurization, it takes a longer time
to achieve complete dissociation. It takes less time to complete
hydrate decomposition by thermal stimulation, however, energy
conservation requires the injection of additional heat in the core.
Therefore, a method combining depressurization and thermal
stimulation has been proposed for hydrate dissociation, for which
a correct choice of the injection temperature and outlet pressure
requires the study of heat transfer during its application.
3.1. The role of temperature stimulation
In simulations of the combined method, the curves of cumulative gas production in real-time increased with increasing temperature, with the increased gas production rates resulting in a faster
complete reaction (Fig. 2). The same was observed in the curves of

Y. Song et al. / International Journal of Heat and Mass Transfer 92 (2016) 766773

769

Fig. 5. (a) Gas generation rate and (b) cumulative gas production over time for different initial water saturations (Sw): 0.28, 0.35, and 0.42.

Fig. 6. (a) Gas generation rate and (b) cumulative gas production over time for different core thermal conductivities (ks): 1.5 and 8.0 W/(mK).

cumulative gas production using only thermal stimulation (Fig. 2).


In contrast, when comparing the cumulative gas production by
thermal stimulation or using the combined method at equal temperatures, the production was always higher with the combined
method. However, this difference decreased with increasing temperature. The combined method promoted the exploitation rate
but its effect gradually diminished as the temperature increased.
This was likely caused by a decrease in the outlet pressure, which

induced a pressure drop in the core. The core pressure reduction


decreased the phase equilibrium temperature and caused
depressurization, which is advantageous for hydrate dissociation.
Moreover, the heat transferred to the core was insufficient at low
stimulation temperature. As a result, a reduction in the outlet pressure strongly improved gas production. Concluding, a suitable
stimulation temperature and outlet pressure must be chosen when
implementing the combined method.

770

Y. Song et al. / International Journal of Heat and Mass Transfer 92 (2016) 766773

Fig. 7. Simulated hydrate saturation (Sh) after 3 min for core thermal conductivities (ks) of (a) 1.5 and (b) 8 W/(mK). Simulated temperature (T) distribution after 3 min for ks
of (c) 1.5 and (d) 8 W/(mK).

3.2. The role of sensible heat


Rapid hydrate dissociation occurred in the first 5 min, after
which the gas generation rate decreased until the end of the
simulation (Fig. 3(a)). Because of the heat flux from the cap- and
base- sediments, the heat absorbed from the periphery was the
same. Therefore, the cumulative gas production was similar for

all cases, i.e., the differences in Cps had little influence on the ultimate cumulative gas production (Fig. 3(b)). In contrast, the gas
production rates were clearly different, with a low Cps resulting
in rapid gas production in the initial dissociation stage (<3 min),
remaining stable for some time, and then quickly decreasing. Conversely, in the high Cps simulation, it took longer to reach the maximum gas generation rate, which remained maximized only for a

Y. Song et al. / International Journal of Heat and Mass Transfer 92 (2016) 766773

771

Fig. 8. Simulated hydrate saturation (Sh) distribution after 18 min for core thermal conductivities (ks) of (a) 1.5 and (b) 8 W/(mK). Simulated temperature (T) distribution
after 18 min for ks of (c) 1.5 and (d) 8 W/(mK).

short time before decreasing rapidly. Therefore, although porous


media with a lower Cps contributed with less sensible heat, the heat
consumed to reach the equilibrium temperature was much less.
Fig. 4 shows the temperature distributions for different Cps in
3 min after the start of the fast reaction. At this point, the cumulative gas production reached 24.41%, 15.05% and 11.56% of the total
for a Cps of 0, 0.8 and 1.6 kJ/(kgK), respectively. The temperature in

Fig. 4(a) was the highest among the three cases, while in Fig. 4(b) it
was generally higher than in Fig. 4(c). The area of low temperature
was largest in Fig. 4(c). These results suggest that increasing the Cps
of the hydrate sediment resulted in a need for more heat to attain
an equilibrium temperature, thereby impeding dissociation. The
temperature changes during dissociation of the hydrate sediments
differed between using depressurization [17] and the combined

772

Y. Song et al. / International Journal of Heat and Mass Transfer 92 (2016) 766773

Fig. 9. (a) Gas generation rate and (b) cumulative gas production over time for different relative permeabilities of gas (krg) and water (krw), for the basic case (case a,
convection heat transfer of water (nw) = 4 and convection heat transfer of gas (ng) = 2), low relative gas permeability (case b, nw = 4 and ng = 4), and high relative water
permeability (case c, nw = 2 and ng = 2).

method, with a temperature decrease with depressurization and


an increase with the combined method. In summary, the effect of
the Cps of the porous media on gas exploitation differed between
the depressurization and the combined method.
Fig. 5 compares the effect of the water content on the gas generation rate and cumulative gas production for different initial Sw.
Only minor differences in production were observed for the three
cases, suggesting that the influence of water content on hydrate
exploitation was weak. In the first 5 min, a low Sw caused a faster
dissociation compared with the other two cases. The gas generation rate for the low Sw reached its maximum first, but was lower
than for the other cases. Nevertheless, the dissociation was completed at the same time for all cases. Concluding, a lower initial
Sw reduced the heat needed to increase the water temperature
and increased the heat from the periphery for hydrate decomposition. Therefore, sediments with lower initial Sw rapidly started to
dissociate. Since less hydrate remained for decomposition, the
maximum rate attained was lower.
When implementing the combined method, a higher Cps resulted
in a relatively higher inhibition of the gas generation rate. Moreover, although the higher initial water content also inhibited gas
generation rate, its effects on the exploitation were weak. The effect
of the sensible heat on the exploitation using the combined method
was similar to the thermal stimulation [18]. As the combined
method increased the generation rate and shortened the reaction
time, a combination of depressurization and thermal injection
proved to be a useful and practical method for gas production from
hydrate-bearing sediments.
3.3. The role of conductive heat flow
At the initial stage of hydrate dissociation, the gas generation
rate and cumulative production were clearly higher when ks was
1.5 W/(mK) compared to ks = 8 W/(mK) (Fig. 6). However, the
reaction with the higher ks was completed faster. In the first

15 min, the cumulative gas production with a lower ks was higher


than with the higher value, with the two curves intersecting after
16 min. From this point onward, the cumulative production was
higher for the higher ks. At the end of the reaction, the two lines
coincided. To analyze the origin of these transitions, the Sh and
temperature distributions were used to predict the influence of
ks on hydrate dissociation with the combined method.
Hydrate decomposed mostly near the core wall, and less in the
core center, as shown in Fig. 7. The Sh distribution for a lower ks
was lower overall after 3 min, while the area with lower Sh was larger (Fig. 7(a)), explaining why the cumulative gas production was
higher for a lower ks (Fig. 6(b)). Regarding the temperature, a
higher ks resulted in a more uniform distribution and higher average value (Fig. 7(d)), whereas the simulation with the lower ks predicted a higher temperature gradient and lower temperature in the
core center (Fig. 7(c)). This resulted from the better and furtherinto-the-core heat transfer porous media with higher ks increasing
the temperature of the core center and requiring less additional
heat for hydrate decomposition.
The point at which the two curves of cumulative gas production
intersected was analyzed in Fig. 8. Fig. 8(b) shows an almost completed hydrate decomposition in the core, while Fig. 8(a) shows a
complete dissociation near the core wall but not in the center after
18 min. From this point on, at a lower ks, hydrates decomposed
slower, unlike for a higher ks, with hydrates dissociating and reaching exhaustion faster. In addition, the radial temperature distribution exhibited a higher gradient (Fig. 8(c)), suggesting limited heat
transfer inside the core. The entire temperature distribution was
relatively uniform with a high average value (Fig. 8(d)), explaining
the uniform hydrate distribution from Fig. 8(b). This uniform distribution and relatively low average value of Sh illustrated why,
after 18 min, the cumulative gas production was higher for the
higher ks.
The effects of ks on the hydrate dissociation using depressurization [17], thermal stimulation [18], or the combined method were

Y. Song et al. / International Journal of Heat and Mass Transfer 92 (2016) 766773

similar. A high ks had a partial inhibitory effect and did not always
promote dissociation, especially after its initial stage. With a suitable injection temperature and outlet pressure, the combined
method increases the hydrate decomposition in porous media
when the heat conductivity coefficient is high.
3.4. The role of convective heat transfer
According to Fig. 9, the krw increased for the high relative water
permeability case (case c) compared to the basic conditions (case
a), showing that the influence of water flow on gas production
was weak, with subtle differences between cases a and c. Increasing the krg accelerated the gas flow. Thus, for case a, the gas in the
dissociation region flowed faster than for the low krg conditions
(case b). The convective heat transfer was slightly more affected
by the krg than the krw in hydrate decomposition using the combined method. However, the krg and krw influence in the process
was minor, suggesting a weak effect of gas and water fluxes on
heat exchange during hydrate decomposition.
For hydrate dissociation by depressurization [17], the convective heat transfer from the gas flow was more efficient than from
the water flow for an increasing dissociation. Moreover, a high
krg implied a reduction in the gas resistance to heat transfer in
the inner core. However, when using thermal stimulation [18] or
the combined method, the ng and nw had little influence on hydrate
dissociation.
4. Conclusions
Based on previous studies of hydrate exploitation by depressurization and thermal stimulation, we extended our investigation to
gas production from hydrate-bearing sediments using a method
combining depressurization and thermal stimulation. The roles
on gas production of the injection temperature, sensible heat, conductive heat and convective heat were analyzed.
We concluded that implementation of the combined method
had a positive effect on the gas production rate compared with
using only thermal stimulation, which decreased with increasing
temperature. Decreasing the outlet pressure was advantageous
for achieving a pressure drop and reducing the core pressure, leading to a decrease in the phase equilibrium temperature and
hydrate decomposition. Moreover, a lower outlet pressure supported gas generation as insufficient heat was transferred to the
core at lower temperature. When implementing the combined
method, a suitable stimulation temperature and outlet pressure
must be chosen.
Increasing the Cps of methane-hydrate-bearing porous media
inhibited the gas production rate using the combined method.
Methane-hydrate-bearing porous media with a higher Cps required
more heat injection from the surrounding environment to improve
the core temperature instead of dissociating hydrates. Similarly,
higher initial water contents inhibited the gas production rates,
even though the effect was relatively weak.
At the initial stage of dissociation using the combined method, a
higher ks negatively affected hydrate dissociation. However, as the
reaction progressed, decomposition accelerated and was completed. In contrast, gas production at a lower ks proceeded at a slow
and steady rate, taking longer to complete the reaction.
With the combined method, hydrate dissociation was relatively
insensitive to heat convection in the water or methane gas.
The combined method promoted the gas production rate and
shortened the dissociation time compared with the single depressurization or thermal stimulation. However, the ultimate cumulative gas production for different injection temperatures, sensible
heat, conductive heat flow, and convective heat transfer conditions

773

was nearly identical for all methods. Thus, heat transfer had little
impact on the cumulative gas production.
This work has improved the understanding of hydrate dissociation, gas generation rates, and cumulative gas production when
using the combined method. To further improve gas production
rates and energy efficiency, future research should focus on the
role of pressure and temperature in the suppression of ice and secondary hydrate formation.
References
[1] Y. Song, C. Cheng, J. Zhao, Z. Zhu, W. Liu, M. Yang, K. Xue, Evaluation of gas
production from methane hydrates using depressurization, thermal
stimulation and combined methods, Appl. Energy 145 (2015) 265277.
[2] X. Li, B. Yang, Y. Zhang, G. Li, L. Duan, Y. Wang, Z. Chen, N. Huang, H. Wu,
Experimental investigation into gas production from methane hydrate in
sediment by depressurization in a novel pilot-scale hydrate simulator, Appl.
Energy 93 (2012) 722732.
[3] B. Li, X. Li, G. Li, J. Feng, Y. Wang, Depressurization induced gas production
from hydrate deposits with low gas saturation in a pilot-scale hydrate
simulator, Appl. Energy 129 (2014) 274286.
[4] S. Falser, A.C. Palmer, K.B. Cheong, T.T. Soon, Temperature increase during the
depressurization of partially hydrate-saturated formations within the stability
region, Energy Fuels 27 (2) (2013) 796803.
[5] X. Ruan, Y. Song, H. Liang, M. Yang, B. Dou, Numerical simulation of the gas
production behavior of hydrate dissociation by depressurization in hydratebearing porous medium, Energy Fuels 26 (3) (2012) 16811694.
[6] Y. Wang, X. Li, G. Li, Y. Zhang, B. Li, Z. Chen, Experimental investigation into
methane hydrate production during three-dimensional thermal stimulation
with five-spot well system, Appl. Energy 110 (2013) 9097.
[7] M. Selim, E. Sloan, Heat and mass transfer during the dissociation of hydrates
in porous media, AiCHE J. 35 (6) (1989) 10491052.
[8] G.J. Moridis, Y. Seol, T.J. Kneafsey, Studies of Reaction Kinetics of Methane
Hydrate Dissociation in Porous Media, Lawrence Berkeley National Laboratory,
2005.
[9] S. Falser, S. Uchida, A. Palmer, K. Soga, T. Tan, Increased gas production from
hydrates by combining depressurization with heating of the wellbore, Energy
Fuels 26 (10) (2012) 62596267.
[10] G. Moridis, T. Collett, Strategies for Gas Production from Hydrate
Accumulations Under Various Geologic Conditions, Lawrence Berkeley
National Laboratory, 2003.
[11] Y. Liu, M. Strumendo, H. Arastoopour, Simulation of methane production from
hydrates by depressurization and thermal stimulation, Ind. Eng. Chem. Res. 48
(5) (2008) 24512464.
[12] X. Li, M. Chen, M. Zhang, R. Xia, Experimental study of natural gas hydrate
dissociation in porous media by thermal stimulation and depressurization, J.
Exp. Mech. 2 (2011) 015.
[13] Y. Bai, Q. Li, Simulation of gas production from hydrate reservoir by the
combination of warm water flooding and depressurization, Sci. China Technol.
Sci. 53 (9) (2010) 24692476.
[14] Z. Feng, Z. Shen, L. Tang, X. Li, S. Fan, Q. Li, Experimental and numerical studies
of natural gas hydrate dissociation by depressurization in different scale
hydrate reservoirs, J. Chem. Ind. Eng. 58 (6) (2007) 1548.
[15] G. Li, L. Tang, C. Huang, Z. Feng, S. Fan, Thermodynamic evaluation of hot brine
stimulation for natural gas hydrate dissociation, J. Chem. Ind. Eng. 57 (9)
(2006) 2033.
[16] L.G. Tang, R. Xiao, C. Huang, Z. Feng, S.S. Fan, Experimental investigation of
production behavior of gas hydrate under thermal stimulation in
unconsolidated sediment, Energy Fuels 19 (6) (2005) 24022407.
[17] J. Zhao, D. Liu, M. Yang, Y. Song, Analysis of heat transfer effects on gas
production from methane hydrate by depressurization, Int. J. Heat Mass
Transfer 77 (2014) 529541.
[18] J. Zhao, J. Wang, W. Liu, Y. Song, Analysis of heat transfer effects on gas
production from methane hydrate by thermal stimulation, Int. J. Heat Mass
Transfer 87 (2015) 145150.
[19] J. Zhao, C. Ye, Y. Song, W. Liu, C. Cheng, Y. Liu, Y. Zhang, D. Wang, X. Ruan,
Numerical simulation and analysis of water phase effect on methane hydrate
dissociation by depressurization, Ind. Eng. Chem. Res. 51 (7) (2012) 3108
3118.
[20] H. Liang, Y. Song, Y. Chen, Numerical simulation for laboratory-scale methane
hydrate dissociation by depressurization, Energy Convers. Manage. 51 (10)
(2010) 18831890.
[21] Y. Song, H. Liang, 2-D numerical simulation of natural gas hydrate
decomposition through depressurization by fully implicit method, China
Ocean Eng. 23 (3) (2009).
[22] Y. Masuda, Y. Fujinaga, S. Naganawa, K. Fujita, K. Sato, Y. Hayashi, Modeling
and experimental studies on dissociation of methane gas hydrates in berea
sandstone cores, in: 3rd International Conference on Gas Hydrates, Salt Lake
City, Utah, (1999), 1822.
[23] R.M. Butler, Thermal Recovery of Oil and Bitumen, Prentice Hall, 1991.

You might also like