Open access peer-reviewed chapter

Direct Numerical Simulation of Hydrate Dissociation in Homogeneous Porous Media by Applying CFD Method: One Example of CO2 Hydrate

Written By

Wu-Yang Sean

Submitted: 10 November 2017 Reviewed: 05 February 2018 Published: 19 September 2018

DOI: 10.5772/intechopen.74874

Chapter metrics overview

704 Chapter Downloads

View Full Metrics

Abstract

Computational fluid method (CFD) is popular in either large-scale or meso-scale simulations. One example is to establish a new pore-scale (m~μm) model of laboratory-scale sediment samples for estimating the dissociation rate of synthesized CO2 hydrate (CO2H) reported by Jeong. It is assumed that CO2H formed homogeneously in spherical pellets. In the bulk flow, concentration and temperature of liquid CO2 in water flow was analyzed by CFD method under high-pressure state. Finite volume method (FVM) were applied in a face-centered packing in unstructured mesh. At the surface of hydrate, a dissociation model has been employed. Surface mass and heat transfer between hydrate and water are both visualized. The initial temperature 253.15K of CO2H pellets dissociated due to ambient warm water flow of 276.15 and 282.15K and fugacity variation, ex. 2.01 and 1.23 MPa. Three tentative cases with porosity 74, 66, and 49% are individually simulated in this study. In the calculation, periodic conditions are imposed at each surface of packing. Numerical results of this work show good agreement with Nihous’ model. Kinetic modeling by using 3D unstructured mesh and CFD scheme seems a simple tool, and could be easily extended to determine complex phenomena coupled with momentum, mass and heat transfer in the sediment samples.

Keywords

  • heat and mass transfer
  • finite volume method (FVM)
  • computational fluid dynamics (CFD)
  • pore-scale flow

1. Introduction

Computational fluid method (CFD) is popular in either large-scale or meso-scale simulations. One example is to establish a new pore-scale (m~μm) model of laboratory-scale sediment samples for estimating the dissociation rate of synthesized CO2 hydrate (CO2H) reported by [1]. To decrease the CO2 concentration in the air, carbon dioxide capture and storage (CCS) is regarded to be an effective way. One concept of CCS is to store CO2 in gas hydrate in sub-seabed geological formation, as was illustrated by [6]. Besides, many studies about the formation and dissociation of CO2 hydrate (CO2H) while stored in the deep ocean or geologic sediment have been introduced. In particular, flow and transport in sediment is multidisciplinary science including the recovery of oil, groundwater hydrology and CO2 sequestration. It reported the measurements of the dissociation rate of well-characterized, laboratory-synthesized carbon dioxide hydrates in an open-ocean seafloor [5]. The pore effect in the phase equilibrium mainly due to the water activity change was discussed in [7]. The reactive transport at the pore-scale to estimate realistic reaction rates in natural sediments was discussed in [3]. This result can be used to inform continuum scale models and analyze the processes that lead to rate discrepancies in field applications. Pore-scale model is applied to examine engineered fluids [4]. Unstructured mesh is well suited to pore-scale modeling because of adaptive sizing of target unit with high mesh resolution and the ability to handle complicated geometries [17, 18]. Particularly, it can easily be coupled with computational fluid dynamics (CFD) methods, such as finite volume method (FVM) or finite element method (FEM). Unstructured tetrahedral mesh used to define the pore structure is discussed in [19]. Another case includes a numerical simulation of laminar flow based on FVM with unstructured meshes was used to solve the incompressible, steady Navier-Stokes equations through a cluster of metal idealized pores by [20].

The objective of this work is to develop a new pore-scale model for estimating the dissociation rate of CO2H in homogeneous porous media. To cooperate with molecular simulation and field-scale simulators, we aim at establishing pore-scale modeling to analyze the simultaneous kinetic process of CO2H dissociation due to non-equilibrium states. Major assumptions in this study are listed as below:

  1. Only dissociation occurred at the surface, no any formation occurred immediately with dissociation.

  2. CO2 dissociated at the surface is assumed to be dissolved into liquid water totally without considering the gas nucleated.

  3. The surface structure does not collapse with the dissociation of CO2H at the surface of pellets.

  4. Homogeneous face-centered packing of multi-CO2H pellets.

  5. Single phase flow coupled with mass, heat, and momentum transfers.

Advertisement

2. Dissociation modeling at the surface

In this study, the dissociation flux (F1) is assumed to be proportional to the driving force, the free energy difference (μ) introduced by [6], presented as

F1=kblRTlnCHsolCIE1

where kbl is the rate constant mol2J1s1m2 of dissociation. According to [21], kbl is listed as below:

kbl=exp(11,729T+26.398)E2

where CHsol is the mole fraction of CO2 in the aqueous solution at equilibrium state with hydrate, and CI means surface concentration in the ambient aqueous solution at the surface of the hydrate CI.

Advertisement

3. Basic transport equations

Flow in the porous media around CO2H is governed by the continuity and the Navier-Stoke’s equations. The advection-diffusion equations of non-conservative type for mass concentration C and temperature T are also considered.

u=0E3
ut+uu=P+1Reu+uT+ρwFn2gE4
Ct+uC=1ReSc2CE5
Tt+uT=1RePr2TE6

where the viscosity, diffusivity, and thermal conductivity of pure water are included in dimensionless parameters such as the Reynolds number, the Schmit number, and the Prandtl number, which are interpolated as functions of temperature and are updated at every computational time step as summarized in Table 1. U and d (=0.001 m) are the velocity of inflow and diameter of hydrate pellet.

No.FunctionDefinition
1D=7.4×1012φMB1/2TηLVA0.6D: diffusion coefficient of CO2 in water
φ(=2.6): association parameter for the solvent water
MB=18gmol1: molecular weight of water
VA=3.4×105m3mol1: molar volume of CO2
ηL[mPas]: viscosity of water
2νL=8.8286×1010T25.3886×107T+8.314×105νLms2: kinematic viscosity
3λL=487.85lnT2173.8λLWKm1: heat conductivity of water
4αL=λLρWCpαLms2: the thermal diffusivity of aqueous phase
ρW=997.1kgm3: density of water
Cp=4,180JkgK: isobaric specific heat, quoted from “Chemical Engineering Handbook”, Japan (1985)
5Peq=expα+βT×103α=44.580 and β = − 10246.28
6CHsol=aexpbP×106+1.321×104T2.292×1021.8×105CHsolmolem3: solubility of hydrate (275.15 K < T < 281.15 K)
a=0.0016T273.150.9211
b=0.0199logT273.15+0.0942
by Aya et al. [10], Yang et al. [11], and Servio and Englezos [12]

Table 1.

Parameters used in this study.

Advertisement

4. Mass transfer

To rewrite Eq. (1), the flux at the surface of the hydrate can be discretized as

kblRTlnCHsolCI=DC=DCIChIE7

where CI is the varying surface concentration calculated locally at each surface cell, C is the centroid concentration, and hI is the thickness of centroid surface cell, as shown in Figure 1.

Figure 1.

Schematic image of discretized surface concentration.

Advertisement

5. Heat transfer

The equation of energy balance at the surface of CO2H is given by

Q̇H+λHTH=λLTLE8

where Q̇H (=HLF1, where HL is the latent heat of hydrate dissociation) is the dissociation heat transferred to the CO2H, λH is the thermal conductivity of hydrate. Dissociation heat per mole of hydrate, HL, is interpolated from [2] as

HL=207,917530.97×TIE9

where TI is the surface temperature. Then, we have

TI=λLhHTL+λHhLTH207,917FcalhLhHλLhH+λHhL530.97FcalhLhHE10

where TL and TH are the temperatures defined at the centroids cell in the aqueous phase and solid hydrate, respectively; hL and hH are half widths of centroid in the aqueous phase and solid hydrate, respectively. Besides, the temperature in the pellet is calculated by using the heat conductivity equation.

Tt=αH2Tx2E11

where αH=λHρHCp is the thermal diffusivity of CO2H. These relative properties of CO2H are quoted from [9], the thermal diffusivity of aqueous phase (αH) of 1.38×107ms2, the heat capacity of hydrate (Cp) of 2080.0JK1kg1, and the thermal conductivity (λH) of 0.324WK1m1. The density of CO2H (ρH) is given as 1116.8kgm3.

Advertisement

6. Computational conditions

Two types of cells, tetrahedrons and triangular prisms, are applied in the present unstructured grid system, as introduced in Figure 2. In detailed, the surface of hydrate uses prism. Both the flow field and inside the pellet are tetrahedral meshes. Upward is the inflow where initially the uniform velocity profile is adopted. Prism mesh and no-slip condition are imposed at the surface of the pellet. To analyze more detailed mass and heat transfer simulatneously, one cell-layer of the prisms that attached to the CO2H surface is divided into at least five very thin layers as referred in [8] for high Prandtl or Schmidt number. The basic parameters of computation are denoted in Table 1. The initial values of dimensionless parameters are listed in Table 2 at the temperatures from 276.15 to 283.15 K. Reynolds number, Schmidt number, and Prandtl number function of the temperature or pressure are listed in Table 2. The minimum grid size of this computational model is listed in Table 3. Lm, Lc, and LT are the applied mesh thicknesses. δm, δc, and δT are the thickness of momentum, concentration, and thermal boundary layers, respectively. The relationship between δm, δc, and δT quoted from the theory of flat plate boundary layer is listed below:

δm=5.48Red2E12
δc=δm1.026·Sc1/3E13
δT=δcPr1/3E14

Figure 2.

Description of mesh in unstructured grid system of 67,104 cells. (a) Overview, (b) surface of CO2H, (c) water, and (d) pellets of CO2H.

CaseReynold numberPrandtl numberFroude numberSchmidt numberPorous ratioTemperature of water (K)Fugacity of equilibrium (MPa)Fugacity (MPa)
150100.02375574%282.153.893
212880276.151.771
31075566%282.153.893
412880276.151.771
51075549%282.153.893
612880276.151.771
7100100.046755282.153.893
81350.062

Table 2.

Calculation conditions of this work.

Cell number (porosity)Sc*δm*δc&T*LmVTL numberLc&T
53,440 (49%)7552.358E-042.252E-051.0E-0552.0E-06
67,104 (66%)7552.345E-032.252E-046.0E-0551.2E-05
77,432 (74%)7552.594E-032.276E-046.0E-0551.2E-05

Table 3.

The thicknesses of boundary layers, δm and δc&T, and grid sizes, Lm and Lc&T (unit: meter).

Values are quoted from “Chemical Engineering Handbook”, Japan (1985)


To follow [22] of Eq. (14), the boundary layer’s thickness for temperature, δT is assumed as the same size as that for mass concentration, δc. For the initial temperature of the CO2H pellet, Tini is assumed as a constant value of 253.15 K.

Advertisement

7. Verification

The in-house code originally developed by [7] has been applied to determine the intrinsic dissociation rate of methane hydrate. The numerical results verified by experimental results are successfully used in calculating one pellet of hydrate in a slow flow rate of high pressure without considering the collapse of hydrate and the nucleation of bubbles referring to [6, 20] as well.

Advertisement

8. Results of case study

In this study, cases with porosity of 74, 66, and 49 are individually discretized as face-centered unstructured packing of hydrate in sediment. CO2H pellets with initial temperature of 253.15 K dissociate due to the variation of driving force, ex. 0.89 and 0.77 MPa, under thermal stimulation of ambient warm water, ex. 282.15 and 276.15 K. Comparative small driving forces selected here is due to the assumption of no surface’s collapse. Computational conditions are listed in Table 2. Result of flux at the surface is the converge value as shown in Figure 3. In Figure 4(a)(c) at time 0.16(s) show velocity vector of case 1 in three specific sections. In Figure 5, the distributions of concentration at 0.16 s are presented. Slight CO2 discharges at the surface. Relative temperature distributions are indicated in Figure 6. As time increases, the dissociation heat of CO2 hydrate results in water temperature drop significantly as shown in Figure 7(a)(c). Relative concentration distribution in center section is shown in Figure 8. The heat of water conducts to the solid-side pellet rapidly in few seconds, and slow mass transfer at the surface dominates the dissociation rate rather than fast heat transfer at the surface.

Figure 3.

Converge value of case 1.

Figure 4.

Velocity vector in the vicinity of pellets (case 1: T = 282.15 K, Tini = 253.15 K, Re = 50, Sc = 755, Pr = 10, VTL = 5, and time = 0.16 s). (a) Front section, (b) Center section and (c) Back section.

Figure 5.

Concentration profile in three specific sections of cubic unit (case 1: time = 0.16 s, unit: mole/m2 s). (a) Front section, (b) Center section and (c) Back section.

Figure 6.

Temperature profile in three specific sections of cubic unit (case 1, time = 0.16 s, unit: K). (a) Front section, (b) Center section and (c) Back section.

Figure 7.

Temperature versus time in center section of cubic unit (case 1, time = 0.16 s, 0.27 s, and 0.54 s, unit: K). (a) Front section, (b) Center section and (c) Back section.

Figure 8.

Concentration versus time in center section of cubic unit (case 1, time = 0.16 s, 0.27 s, and 0.54 s, unit: mole/m2 s). (a) Front section, (b) Center section and (c) Back section.

Advertisement

9. Discussion

To follow the modeling illustrated in [14]:

F3=kDfeqfg=kD0expΔERTfeqfgE15

where kD0molPa1s1m2 is the rate constant, fgPa is the fugacity of gaseous CO2, and feq is the fugacity of the quadruple equilibrium. They obtained kD0 and E for CO2H as 1.83×108molPa1s1m2 and 102.88kJmol1, respectively, at temperature and pressure ranging from 274.15 to 281.15 K and from 1.4 to 3.3 MPa. However, new modified value of E, if considered the real case in the ocean quoted from [16], is 96.49kJmol1. The order of Reynolds number based on the size of a particle, about 16 μm, is calculated as 50. Clarke et al. [28] determined the dissociation rate of CO2H by measuring the nucleated methane gas in V-L state [14]. The comparison of three models is listed in Table 4. The results of dissociation flux are summarized in Figure 9. Higher water temperature induces higher dissociation flux at the surface of hydrate. Data correlated by [14] show much lower level than both Nihous’ model and new model. The numerical results in this work marked as new model in Figure 9 show consistent result compared with Nihous’ model. The dissociation flux in various flow rates in cases 5, 7, and 8 are listed in Figure 10. Here, it is noted that porosity is not considered in both Clarke’s and Nihous’ models, and these two models are only function of temperature and fugacity as presented in Eq. (15). The trend of flux becomes saturated in the figure due to the surface dissociation flux that becomes slow due to the fast mass transfer in bulk flow at Reynolds number over 100.

ItemModelingIntrinsic rate of dissociationRef.
Clarke’s ModelEq. (15)KD0 = 1.83 * 108mole/Pa s m; ΔE = 102.89 kJ/mole[14]
Nihous’ ModelEq. (15)KD0 = 1.83 * 108mole/Pa s m; ΔE = 96.49 kJ/mole[16]
New ModelEq. (1)Kbl = exp(−11,729/T + 26,398)[21]

Table 4.

The comparison of three models.

Figure 9.

Results of simulation compared to existing two models.

Figure 10.

Results of flux in cases 5, 7, and 8.

Advertisement

10. Conclusions

The objective of this work is to establish a new pore-scale model for estimating the dissociation rate of CO2H in laboratory-scale sediment samples. It is assumed that CO2H formed homogeneously in spherical pellets. In the bulk flow, concentration and temperature of liquid CO2 in water flow was analyzed by computational fluid dynamics (CFD) method without considering gas nucleation under high-pressure state. In this work, finite volume method (FVM) was applied in a face-centered regular packing in unstructured mesh. At the surface of hydrate, a dissociation model has been employed. Surface mass and heat transfer between hydrate and water are both visualized. The initial temperature 253.15 K of CO2H pellets dissociated due to ambient warm water flow of 276.15 and 282.15 K and fugacity variation, ex. 2.01 and 1.23 MPa. Three tentative cases with porosity 74, 66, and 49% are individually simulated in this study. In the calculation, periodic conditions are imposed at each surface of packing. Additionally, the flux at CO2H’s surfaces is compared to Clarke and Bishnoi [13] and Nihous and Masutani [15] ‘s correlations at Reynolds number of 50. Numerical results of this work show good agreement with Nihous’ model. Kinetic modeling by using 3D unstructured mesh of regular cubic unit and CFD scheme seems to be a simple tool to estimate the dissociation rate of CO2H in laboratory-scale experiments, and could be easily extended to determine complex phenomena coupled with momentum, mass, and heat transfer in the sediment samples.

Acknowledgments

This work was supported by DOIT, Ministry of Science and Technology under contract No. MOST 106-3113-M-002-006. The authors also wish to acknowledge Professor Toru SATO for the valuable advices and guidance.

Nomenclature

Cvolumetric molar concentration of CO2 in the ambient water molm−3
CHvolumetric molar concentration of CO2 in the aqueous solution equilibrated with the stable hydrate phase molm−3
C′volumetric molar concentration of CO2 in water at the centroid of a cell attaching to the hydrate surface molm−3
CIvolumetric molar concentration of CO2 in the ambient aqueous solution at the surface of the hydrate ball molm−3
CXaverage molar volumetric concentration of CO2 in the ambient water flow for a given cross section of water flow molm−3
ddiameter of the CO2 hydrate ball [m]
Ddiffusion coefficient of CO2 in water ms−2
Eactivation energy Jmol−1
Fdissociation rate flux mols−1m−2
feqfugacity of the quadruple equilibrium Pa.
fgfugacity of gaseous CO2 Pa
Gmolar Gibbs free energy Jmol−1
HLlatent heat of hydrate dissociation Jmol−1
hLlength of the water layer attached to the hydrate surface [m]
kD0intrinsic dissociate rate constant based on Clarke-Bishnoi model molPa−1s−1m−2
kbldissociation rate constant based on new model mol2J−1s−1m−2
Lthickness of computational cell [m]
MBmolecular weight of water gmol−1
Pthermodynamic pressure [Pa]
Peqquadruple equilibrium pressure for CO2 hydrate as a function of T [Pa]
Qvolumetric flow rate of the ambient water m3s−1
Q̇Hthe rate at which the latent heat is transferred to the CO2 hydrate by dissociation Jm−2s−1
Rgas constant, 8.314 JK−1mol−1
Tabsolute temperature [K]
TLtemperature at the centroids of a cell in the solid hydrate [K]
THtemperature at the centroids of a cell in the aqueous phase [K]
uvelocity vectors [m/s]
xmole fraction of CO2 [-]
xeqsolubility of CO2 in the aqueous solution in equilibrium with the stable hydrate phase [-]
xImole fraction of CO2 in the aqueous phase at the surface of the hydrate ball [-]
αLthermal diffusivity in the aqueous phase ms−2
αHthermal diffusivity in the hydrate ball ms−2
Δrthickness of the computational cell [m]
δthickness of the boundary layer [m]
Δμchemical potential difference Jmol−1
ρwdensity of the ambient water kgm−3
φthe association parameter for the solvent water
ηLthe viscosity of water [mPas]
VAthe molar volume of CO2 m3mol−1
νLkinematic viscosity of water ms−2
λLheat conductivity of water WK−1m−1
λH and λLthe heat conductivities in the hydrate and water WK−1m−1

References

  1. 1. Jeong S-M, Hsieh L-HC, Huang C-Y, Sean W-Y. Direct numerical simulation of CO2 hydrate dissociation in pore-scale flow by applying CFD method. International Journal of Heat and Mass Transfer. 2017;107:300-306
  2. 2. Andersson V, Kvaerner A, Haines M. Gas hydrates for deep ocean storage of CO2—novel technology for utilising hydrates for transport of CO2. In: Wilson M, Morris T, Gale J, Thambimuthu K, editors. Greenhouse Gas Control Technologies, Volume II. New York: Elsevier; 2005. pp. 1487-1492
  3. 3. Molins S, Trebotich D, Steefel CI, Shen C. An investigation of the effect of pore scale flow on average geochemical reaction rates using direct numerical simulation. Water Resources Research. 2012;48:7453-7460. DOI: 10.1029/2011WR011404
  4. 4. Davison SM, Yoon H, Martinez MJ. Pore scale analysis of the impact of mixing-induced reaction dependent viscosity variations. US Department of Energy Publications. Paper 110. 2012. http://digitalcommons.unl.edu/usdoepub/110
  5. 5. Gregor R, Stephen HK, William BD, Laura AS, Edward TP, John P, Peter GB. Dissolution rates of pure methane hydrate and carbon-dioxide hydrate in undersaturated seawater at 1000-m depth. Geochimica et Cosmochimica Acta. 2004;68(2):285-292
  6. 6. Inui M, Sato T. Economical feasibility study on CO2 sequestration in the form of hydrate under seafloor. Proceedings of 26th International Conference on Offshore Mechanical and Arctic Engineering. 2006;OMAE06-92306:1-10
  7. 7. Uchida T, Ebinuma T, Takeya S, Nagao J, Narita H. Effects of pore sizes on dissociation temperatures and pressures of methane, carbon dioxide, and propane hydrates in porous media. The Journal of Physical Chemistry. B. 2002;106:820-826
  8. 8. Jung RT, Sato T. Numerical simulation of high Schmidt number flow on unstructured hybrid mesh. Journal of Computational Physics. 2004;203:221-249
  9. 9. Sloan ED. Clathrate hydrate of natural gases. New York: Dekker; 1998
  10. 10. Aya I, Yamane K, Nariai H. Solubility of CO2 and density of CO2 hydrate at 30MPa. Energy. 1997;22:263-271
  11. 11. Yang SO, Yang IM, Kim YS, Lee CS. Measurement and prediction of phase equilibria for water + CO2 in hydrate forming condition. Fluid Phase Equilibria. 2000;175:75-89
  12. 12. Servio P, Englezos P. Effect of temperature and pressure on the solubility of carbon dioxide in water in the presence of gas hydrate. Fluid Phase Equilibria. 2001;190:127-134
  13. 13. Clarke MA, Bishnoi PR. Determination of the intrinsic rate constant and activation energy of CO2 gas hydrate decomposition using in-situ particle size analysis. Chemical Engineering Science. 2004;59:2983-2993
  14. 14. Sean W, Sato T, Yamasaki A, Kiyono F. CFD and experimental study on methane hydrate dissociation part II. General cases. AICHE Journal. 2007;53:2148-2160
  15. 15. Nihous GC, Masutani SM. Notes on the dissolution rate of gas hydrates in undersaturated water. Chemical Engineering Science. 2006;61:7827-7830
  16. 16. O'Connor RM, Fredrich JT. Microscale flow modelling in geologic materials. Physics and Chemistry of the Earth, Part A: Solid Earth and Geodesy. 1999;24(7):611-616
  17. 17. Piller M, Nolich M, Favretto S, Radaelli F, Rossi E. Analysis of hydraulic permeability in porous media: From high resolution X-ray tomography to direct numerical simulation. Transport in Porous Media. 2009;80(1):57-78
  18. 18. Sahimi M. Flow and Transport in Porous Media and Fractured Rock. Weinheim: VCG; 1995
  19. 19. Boomsma K, Poulikakos D, Ventikos Y. International Journal of Heat and Fluid Flow. 2003;24:825
  20. 20. Fukumoto A, Sean WY, Sato T, Yamasaki A, Kiyono F. Estimation of dissociation rate constant of CO2 hydrate in water flow. Greenhouse Gases: Science and Technology. 2015;5(2):169-179
  21. 21. Pohlhausen E. Der Wärmeaustausch zwischen festen Körpern und Flüssigkeiten mit kleiner reibung und kleiner Wärmeleitung. Zeitschrift für Angewandte Mathematik und Mechanik. 1921;1:115-121. DOI: 10.1002/zamm.19210010205
  22. 22. Shu SS, Lee MJ. Dynamic behavior of methane hydrates formation and decomposition with a visual high-pressure apparatus. Journal of the Taiwan Institute of Chemical Engineers. 2016;62:1-9

Written By

Wu-Yang Sean

Submitted: 10 November 2017 Reviewed: 05 February 2018 Published: 19 September 2018