Guest Migration Revealed in CO2 Clathrate Hydrates - ACS Publications

Aug 13, 2015 - T. C. Hansen,. § and W. F. Kuhs*,‡. †. Department of Applied Mathematics, Kazan (Volga Region) Federal University, 420008 Kazan, R...
0 downloads 0 Views 657KB Size
Subscriber access provided by UNIV PRINCE EDWARD ISLAND

Article

Guest Migration Revealed in CO2 Clathrate Hydrates Andrey N Salamatin, Andrzej Falenty, Thomas C Hansen, and Werner F Kuhs Energy Fuels, Just Accepted Manuscript • DOI: 10.1021/acs.energyfuels.5b01217 • Publication Date (Web): 13 Aug 2015 Downloaded from http://pubs.acs.org on August 21, 2015

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Energy & Fuels is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1

Energy & Fuels

Guest Migration Revealed in CO2 Clathrate Hydrates A.N. Salamatin1, A. Falenty 2, T.C. Hansen3, and W.F. Kuhs 2,*

2 3 4

1

Dept. of Applied Mathematics, Kazan (Volga Region) Federal University, 420008 Kazan, Russia 2

GZG Abt. Kristallographie, Georg-August-Universität Göttingen, 37077 Göttingen, Germany 3

5

Institut Laue-Langevin (ILL), 71 avenue des Martyrs, 38000 Grenoble, France

6 7

* Werner F. Kuhs, [email protected], tel. +49 551-39-3891, fax. +49 551-39-95-21

8

RECEIVED DATE (to be automatically inserted after your manuscript is accepted if required

9

according to the journal that you are submitting your paper to

10

Abstract

11

The shrinking-core model of the formation of gas hydrates from ice spheres with well-defined

12

geometry gives experimental access to the gas permeation in bulk hydrates which is relevant to their

13

use as energy storage materials, their exploitation from natural resources as well as to their role in

14

flow assurance. Here we report on a new approach to model CO2 clathration experiments in the

15

temperature range from 230 to 272 K. We develop a comprehensive description of the gas

16

permeation based on the diffusion along the network of polyhedral cages, some of them being empty.

17

Following earlier molecular dynamics simulation results, the jump from a cage to one of its empty

18

neighbors is assumed to proceed via a “hole-in-cage-wall” mechanism involving water vacancies in

19

cage walls. The rate-limiting process in the investigated temperature range can be explained by the

20

creation of water-vacancy-interstitial pairs. The gas diffusion leads to a time-dependent cage filling

21

which decreases across the hydrate layer with the distance from the particle surface. The model 1 ACS Paragon Plus Environment

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 28

1

allows a prediction of the time needed for a complete conversion of ice spheres into clathrate as well

2

as the time needed for a full equilibration of the cage fillings. The findings essentially support our

3

earlier results obtained in the framework of a purely phenomenological permeation model in terms of

4

the overall transformation kinetics, yet it provides for the first time insight into the cage equilibration

5

processes. The diffusion of CO2 molecules through bulk hydrate is found to be about three to four

6

times faster in comparison with the CH4 case.

7 8

Keywords: gas hydrate, gas diffusion, hole-in-the-cage mechanism, water vacancies, ice, neutron

9

diffraction, Rietveld refinement

10 11

Introduction

12

Clathrate gas hydrates (GH), nonstoichiometric crystalline compounds of guest-gas molecules built

13

into small and large cages of a metastable framework of water molecules, are a widely spread matter

14

in geosystems and are important in technological applications for energy storage and flow assurance.

15

Different physical scenarios with gas/water mixtures or solutions can result in gas hydrate

16

appearance. Here we focus on the GH formation process which occurs on the ice/water-gas interfaces

17

(i.e. excluding a direct nucleation from liquid). This phenomenon has been studied under laboratory

18

conditions by exposing ice particles (e.g.1-4) or liquid-water (e.g.5-7) to clathrate-forming gas. It is

19

now well established that, in spite of obvious differences between gas-ice and water-gas interfaces,

20

in both cases the gas hydrate formation follows a similar pattern: 1) The ice/water-gas interface is,

21

first, covered by thin hydrate film composed of initially nucleated hydrate patches e.g.2, 4. The initial

22

thickness of this layer is limited to a few µm e.g.2, 8. 2) After this relatively short stage I, the next

23

stage II sets on when the further clathration reaction is maintained and, generally, limited by guest-

24

gas molecules permeation (diffusion) from the environmental gaseous atmosphere through the

25

hydrate layer towards the unreacted ice/water phase. At the same time water molecules move to the

26

outer boundary of the hydrate layer to react there with the ambient gas. What drives the gas transport

27

is the supersaturation of the "gas-ice/water-hydrate" system, i.e. the difference between the chemical

ACS Paragon Plus Environment

2

Page 3 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

1

potential of the gas encaged in GH at the reaction front, at the contact with ice/water phase, and that

2

on the outer GH surface at the contact with the gaseous phase.

3

Powder samples of small spherical ice particles with well defined specific surface turned out to be

4

particularly helpful in accessing individual stages and establishing apparent activation energies for

5

the gas migration through the clathrate lattice. An elaborated description of the initial coating stage

6

of hydrate growth on ice spheres as a stochastic “birth-and-growth” phenomenon (i.e. an interplay of

7

two mechanisms: (a) creation of "two-dimensional" nucleation spots and (b) limited lateral growth of

8

the hydrate patches on the ice surface) was developed and employed for interpretation of the pVT

9

(in-house) and neutron diffraction hydrate formation experiments2 on the basis of the interactive

10

computer system "POWDER-4". With only small changes in resulting predictions of the

11

transformation kinetics, the stage-I sub-model is now improved to represent a more realistic scenario

12

of ice-powder structure development at microscopic randomly uniform hydrate patch nucleation2.

13

As for the stage II modeling, our earlier study2-4 and the software assumed that the gas permeation

14

mechanisms were linked to polydispersity / polycrystallinity of GH build-up (particle/grain

15

boundaries, triple junctions), linear crystallographic defects, and other imperfectness of a crystalline

16

structure. A certain, though small, amount of mobile gas molecules was envisaged as an important

17

bulk property of clathrate hydrates. This pseudo-solute component could be acquired from ambient

18

atmosphere to be transported further by the quasi-stationary permeation (diffusion) process to the

19

ice-to-hydrate transformation front. Among many other works, such an approach was explored in a

20

series of our publications3,

21

modeling efforts10-12 based on formulation of the general diffusion equations in the hydrate layer do

22

not specify the gas permeation process on the microscopic level of hydrate structure and thus remain

23

essentially at the level of approximation spelled out in our earlier work3,

24

attempt to cast the model into a more physical picture of the transport phenomena.

4, 9

, but remained purely phenomenological. Unfortunately, recent

4, 9

. In the following we

25

A review of recent studies and data reconciliation2 showed that one of the most likely transport

26

mechanisms of gas molecules through the crystalline hydrate structure might be the so-called "hole-

27

in-cage-wall" migration realized via water vacancies in cage walls connecting neighboring cages.

28

Small molecules like H2 and Ne appear to move through the GH framework without water vacancies ACS Paragon Plus Environment

3

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 28

1

for jumps between large cages (passing through the 6-membered connecting water rings), while

2

jumps involving a passage through 5-membered rings appear to involve water defects even for such

3

small molecules13. As a consequence, the migration process also involves a transient cage-filling

4

readjustment during the hydrate shell growth and some gas exchange between small and large cages.

5

As for the water interstitial transport from the ice-hydrate interface to the particles’ surface, in

6

accordance with the present-day understanding14, it is not the rate limiting process for the gas hydrate

7

formation from ice. Here we continue our research line and complete the development of the

8

generalized model for the non-equilibrium "hole-in-cage-wall" diffusion15 of the guest-gas molecules

9

through the hydrate layer of structure I (sI) growing around an ice particle; this mechanism has also

10

been called “water-vacancy-assisted”16 and relates to a very dynamic opening of a water ring

11

structure in polyhedral cage walls of a hydrate structure by a temporary creation of a water vacancy

12

permitting the passage of guest molecules. A principal peculiarity of the sI-hydrate build-up related

13

to modeling the diffusive gas transport is that, in contrast to hydrate structure II (sII), the smaller

14

cages (SC-s) in sI are fully surrounded by larger cages (LC-s) and do not have direct "common-wall"

15

contacts with each other. Thus, in addition to the overall cage occupancy transition, a special

16

emphasis must also be paid to the gas-mass exchange between LC-s and SC-s. We, further, include

17

the diffusion theory into the macroscopic description of the single-gas sI-hydrate formation from ice

18

powders. The developed, more comprehensive and more physical (but still phenomenological) model

19

is implemented as a modified computational algorithm in the next POWDER-5 version of the

20

POWDER-software series and is applied to interpret CO2-hydrate formation experiments.

21 22

Experimental

23

CO2-hydrate was grown from well-defined spherical ice powders prepared in a standardized

24

procedure3, 4, 9 by spraying deuterated water into liq. N2 starting from 99.9 % purity D2O. This was

25

done in a sealed glove-box under N2 saturated atmosphere in order to avoid possible contamination

26

(and/or dilution) with atmospheric water. Large ice particles were removed by sieving under liq. N2

27

through a 350 µm mesh, and the resulting material was stored under liq. N2. Cryo-Scanning electron

28

microscopy revealed a log-normal distribution of ice spheres with a mean diameter of 52 µm. These ACS Paragon Plus Environment

4

Page 5 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

1

ice spheres consist of grown-together individual ice crystals of 5-15 µm in diameter. The mean

2

volume-radius of 42.64 µm has been deduced from the volume distribution of ice spheres.

3

Ice powders were transferred into thin walled Al vials and compacted to a remaining porosity of

4

~30-40 % (estimated from volume/mass ratio). These samples containing ~ 1 g of the starting

5

material were inserted in a pressure cell before CO2 gas was applied with pressures below the

6

stability limit of CO2-hydrate. CO2 pressures were measured with a Ashcroft KXD linear gauge (6

7

MPa). The reaction rig built around the high-flux 2-axis neutron diffractometer D20 at the Institut

8

Laue-Langevin, Grenoble, France17 was used to explore the formation reaction. The excellent

9

penetration power of neutrons, the high flux available on D20 instrument and its simultaneous

10

readout over a 2θ range of 153.6 deg (at λ = 2.418 Å) gave the unique advantages to follow in-situ

11

the nucleation and growth phase as well as the later slow transition to the diffusion limited stage

12

(Fig. S1). The acquisition time was set typically to 30-60 s for the initial and 300 s for later parts of

13

the formation, depending somewhat on the currently observed reaction rate. Temperature control was

14

achieved with a PID regulated “orange” He-flow cryostat operating between 1.7 and 300 K with

15

accuracy better than 0.1 K. The large set of diffraction data was analyzed in an automated fashion

16

using the full pattern Rietveld refinement package GSAS18 (Fig S1). The phase fractions of the ice

17

starting material and the newly formed CO2-hydrate were established with accuracies better than 1

18

wt.%. These results present an average for the complete sample volume of ~ 1 cm3. The profile

19

functions, zero point, lattice constants of GH, atomic positional and displacement parameters for

20

D2O ice Ih and CO2-hydrate were kept fixed once established (Table S1). The freely refined

21

parameters were the lattice constants of ice Ih, the phase fractions, five to six parameters of the

22

background description by Chebyshev polynomials and occupancies of the small and large cages. It

23

should be noted that the retrieval of cage fillings is a challenging task even with good quality

24

diffraction data2-4 and typically requires a number of constraints. Particularly troublesome is the

25

strong correlation between the cage occupancies and the atomic displacements (Uiso) that may lead to

26

significant overestimations of gas occupancies in LC-s and underestimation in SC-s. Here these

27

parameters have been tackled by fixing thermal displacement parameters to values derived from

28

expansivity data19 cross correlated with earlier, high resolution neutron20 and X-ray powder ACS Paragon Plus Environment

5

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 28

1

diffraction21 measurements. Correlations with instrumental parameters (zero point, instrumental

2

profiles and sample position) turned out to be less troublesome and could be safely avoided by

3

careful calibration on a reference material (Na2Ca3Al2F14).

4 5

Model Description

6

Thermodynamic properties of GH

7

Two basic crystallographic GH structures, the Stackelberg Structure I (sI) and Structure II (sII), are

8

distinguished22, depending on the guest-gas nature23. The unit cell characteristic of the sI clathrate

9

water host lattice is summarized below in Table 1 following22, 24, 25. It is important to note that the

10

density of metastable water frameworks in GH, ρw ≈ 44 kmol/m3, is essentially lower than the Ih-ice

11

density ρi = 51 kmol/m3. Thus, structural changes during ice-to-hydrate transformation must proceed

12

with noticeable water mass outflow from the ice reaction front. Hereafter the two types of smaller

13

and larger cavities are referred to by lower indices j =1, 2, respectively. General theoretical

14

considerations22 express the chemical potential µw of the host-water framework in clathrate hydrate

15

crystal via the guest-gas occupancies yj,

µ w = µ w ( p, T ) + RT [ν 1 ln(1 − y1 ) + ν 2 ln(1 − y 2 )] .

16

(1)

17

Here µ w is the chemical potential of the empty metastable water host lattice, being a function of

18

external pressure p and Kelvin temperature T within the GH stability region; νj is the number of

19

cages of the j-th type per one water molecule; R is the gas constant. In thermodynamic equilibrium

20

with the ambient gaseous atmosphere characterized at the given p- and T- conditions by the gas

21

fugacity fa, one has

22

fa =

1 y1a 1 y 2a , = C1 1 − y1a C 2 1 − y 2 a

(2)

23

where Cj are the temperature-dependent Langmuir constants of the gas in the SC and LC hydrate

24

cavities and yja are the corresponding equilibrium cage occupancies, j = 1, 2. In this work C1 and C2

25

are derived from the Langmuir isotherm fitted to cage occupancies of CO2-hydrate calculated with

ACS Paragon Plus Environment

6

Page 7 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

1

CSMGem23. Accordingly, the chemical potentials µj of the gas molecules encaged in SC-s and LC-s

2

can be expressed via their apparent fugacities fj or occupancies yj as

3

 1 yj  C j 1− y j

µ j = RT ln f j = RT ln

  , j = 1, 2.  

(3)

4

We also introduce the guest-gas density ρg in hydrate related to the molar concentration of the

5

encaged gas cg counted per a unit molar mass of host-water molecules,

6 7 8

ρ g = ρ wcg ,

c g = ν 1 y1 + ν 2 y 2 ,

(4)

Inversely, the GH-properties (1)-(4) can be rewritten in terms of the apparent gas fugacities with yj =

Cj f j 1+ C j f j

, j = 1, 2.

(5)

9 10

General description of guest-gas diffusion in sI-hydrate

11

Various natural phenomena and laboratory observations e.g.4,

25, 26

12

molecules can penetrate ("diffuse") through bulk hydrate barriers without direct recrystallization

13

and/or reformation of the lattice cages. As reviewed and discussed2, in recent years molecular

14

modeling techniques have revealed plausible mechanisms of the guest-gas permeation in single-

15

crystalline hydrates. The principal conclusion is that, most probably, the gas diffusion is assisted by

16

the presence of empty cages and water vacancies in the cage walls, i.e. is the so-called "hole-in-cage-

17

wall" permeation. For instance, in CO2-hydrate of structure I, around 30 % of small cages and a few

18

percent of large cages are empty at the typical conditions of laboratory experiments. However, small

19

cavities in sI-hydrate are not directly linked to each other. Moreover, the activation energy to hop

20

from SC-s to LC-s was found27 to be 1.5-2 times higher than that of hops between LC-s in the

21

presence of water vacancies in the hydrate framework. Therefore, one can assume that the long-range

22

transport in sI-hydrate mainly proceeds through LC-s, while SC-s work rather as sinks or sources of

23

guest-gas molecules, facilitating the gas transport by evacuating guests from LC-s, although without

24

direct participation as the permeation paths in the diffusion process.

demonstrate that the guest-gas

25

Concerning the transport of water molecules involved in the clathration process3, water interstitials

26

are suggested as a viable mechanism. Interstitial water transport is the likely mechanism for proton ACS Paragon Plus Environment

7

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 28

1

transport in ice Ih28 and this is also a possible mechanism for hydrates29 to be even more effective

2

than a vacancy mechanism. The water mobility itself is not thought to be a rate-limiting process for

3

gas permeation14 – rather, at high temperatures, it is the creation of water vacancy-interstitial pairs15

4

that controls the guest transport rates. As shown in MD simulations14, interstitial water molecules

5

can move easily in the host framework and can interchange with host water molecules in various

6

ways and most easily when the cages are empty. Thus, we might well model the inward migration of

7

gas molecules via a hole-in-cage-wall mechanism and assume the outward migration of water

8

molecules to take place via the interstitial mechanism.

9

Let us consider the hole-in-cage-wall migration of gas molecules realized via water vacancies

10

connecting neighboring LC cavities in the crystalline hydrate structure I. This mechanism obviously

11

involves the guest-gas redistribution in the GH bulk, including gas exchange between LC-s and SC-s

12

and the cage filling transient readjustment. In accordance with the general concepts of non-

13

equilibrium thermodynamics30,

14

hydrate lattice we write after32

31

, for mass flux qg of gas molecules through larger cages in sI-

q g = − λ g ∇µ 2 ,

15

(6)

16

where λg is the mass transport (phenomenological) coefficient and ∇ is the differential Nabla-

17

operator. Hereinafter we conventionally assume that the λg-factor in Eq. (6), i.e. gas-mass transport

18

via LC-s, is proportional to the molar concentration of the guest-gas molecules in these cavities.

19

Furthermore, in contrast with the classical diffusion theory e.g.31, the potential gas mobility ϑg,

20

depending on water vacancy concentration in GH, must be additionally reduced by the probability for

21

gas molecules to meet an empty large cage. Let us also note that the probability for the guest

22

molecules to move between LC-s via intermediate jump to/from SC-s is, at least, one order of

23

magnitude less and, thus, is neglected in our further considerations. Consequently, in accordance

24

with the hole-in-cage-wall diffusion scenario, the following representation for the phenomenological

25

coefficient can be envisaged:

26 27

λ g = ρ wν 2 y 2 (1 − y 2 )ϑ g .

(7)

Combining Eqs. (3), (6), and (7), one finally arrives at the Fick's law ACS Paragon Plus Environment

8

Page 9 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

q g = − ρ wν 2 D g ∇y 2 ,

1

(8)

2

where Dg = RTϑg is, by definition, the diffusion coefficient in sI-hydrate through the LC cavities. We

3

note here that the diffusion constant is not assumed to have an explicit dependency on the cage

4

fillings; this is justified by molecular dynamics computer simulations for CH4 hydrates16, where -

5

within our experimentally established range of cage fillings - changes of only a few % are expected.

6

The next step in modeling the gas-mass transport through the newly-formed non-equilibrium

7

hydrate bulk is to describe the gas exchange between the smaller and larger cages. This is a

8

relaxation phenomenon which reveals itself as a sink/source of gas molecules in GH cavities of

9

different types, depending on their initial filling (i.e. chemical potentials of guest molecules). Hence,

10

we introduce the flux σg of gas molecules from LC-s to SC-s in a unit of GH (water framework)

11

volume per a unit of time. Formally one can present σg in two different ways, as

σ g = λ 21 ( µ 2 − µ1 ) or σ g = −λ12 ( µ1 − µ 2 ) ,

12

(9a)

13

where λ21 and λ12 are the mass transport coefficients of the gas molecules from LC-s to SC-s at

14

µ2 > µ1 and from SC-s to LC-s at µ1 > µ2, respectively.

15

Similarly to Eq. (7), we write

λ21 = ρ wν 1ν 2

16

C1 y 2 (1 − y1 ) C y (1 − y 2 ) ϑ21 and λ12 = ρ wν 1ν 2 2 1 ϑ12 . C1 + C 2 C1 + C 2

(9b)

17

Here ϑ21 and ϑ12 are the analogues of potential gas mobilities proportional to hopping frequencies of

18

gas molecules between cages of different types. In proximity to thermodynamic equilibrium

19

ϑ21 ≈ ϑ12 .

20 21 22

23 24

Accordingly, with the use of Eqs. (3), the driving force of the mass exchange process in Eqs. (9a) becomes

µ j − µ k = RT ln

 C j y k (1 − y j )   , ≈ RT 1 −  C y (1 − y )  fk k j k  

fj

and together with Eqs. (9b) this results in the following form of the mass flux

σg =

ρ wν 1ν 2 χ g C1 + C 2

[C1 y 2 (1 − y1 ) − C 2 y1 (1 − y 2 )] ,

ACS Paragon Plus Environment

(10)

9

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1

Page 10 of 28

where χg = RTϑ21 ≈ RTϑ12 is the mass exchange coefficient.

2

Finally, the mass conservation laws for the guest gas molecules occluded in smaller and larger

3

cavities of non-equilibrium gas hydrate combined with constitutive equations (8) and (10) yield the

4

generalized diffusion model

5

ν 2χg ∂y1 [C1 y 2 (1 − y1 ) − C 2 y1 (1 − y 2 )] , = ∂t C1 + C 2

∂y 2 ν 1 ∂y1 + = ∇ ⋅ ( D g ∇y 2 ) . ∂t ν 2 ∂t

(11)

6 Here t is the time. 7 8

Ice-sample structure and initial stage I

9

We consider a starting ice powder as a random dense packing of spherical ice grains with the log-

10

normal size distribution characterized by initial mean radius , initial mean-volume radius

11

r i 0 = ri 0

12

follow2, 3 and use the geometrical description of the local ice-powder structure evolution during the

13

GH formation as developed33 for a random dense packing of expanding mono-disperse spheres. As

14

shown by Kuhs et al4, even in case of the polydisperse ice particle ensemble, the mono-size

15

approximation of the ice-to-hydrate transformation process remains valid up to the reaction (ice-to-

16

hydrate transformation) degree α of 40-50 %. However, the apparent initial radius of ice spheres is

17

now defined so as to represent the specific surface area (SSA) of the starting material that is the

18

initial equivalent-surface-area radius ri 0 = (1 + v0 )r i 0 . As a consequence, this will change the

19

reference particle dimensions and will affect the values of the inferable kinetic parameters, resulting

20

in more realistic estimates. On average, locally the external shape of hydrate layers formed on ice

21

particles is represented as a truncated sphere of radius rh. The ice cores shrink, their radius ri

22

decreases, due to the inward growth of the hydrate shells. But, because of the lesser density of water

23

in the hydrate phase, the excess water molecules must be transported to the outward hydrate surface

24

exposed to the ambient gas, and the hydrate layers simultaneously expand into the open space (voids)

25

between the original ice grains3. This process is governed by the hydrate volume expansion factor

26

E = ρ i ρ w − 1 , and its detailed description in terms of the relative geometric characteristics

3 1/ 3

and relative size variance ν02. As before in the POWDER-4 software model, we

2

ACS Paragon Plus Environment

10

Page 11 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

1

Ri = ri/ri0 and Rh = rh/ri0 is summarized for convenience after2,

2

Information).

3

in Appendix A (Supporting

3

The ice-particle surface fraction αS covered by hydrate film and/or occupied by developing

4

contacts of growing hydrate layers during the initial coating stage I is one of the principal overall ice-

5

to-hydrate conversion characteristics. In accordance with34, let δ0 be the thickness of the surface ice

6

layer converted to the initial hydrate shell. Further, we follow the JMAK kinetic theory35 of

7

recrystallization and employ the improved modeling approach2 presented in Appendix B (Supporting

8

Information) to simulate the hydrate film formation on ice spheres. As before, we introduce the

9

general notion of the ice-sphere coating rate ΩS, which is, by definition, the fraction of the free

10

(exposed to the ambient gas) ice surface which becomes covered by the newly nucleated hydrate

11

patches during a unit of time. The Vandermeer-Rath microstructural path methodology35 assumes

12

that the apparent radius of a hydrate patch develops with its age τ as 2Gτ m/2, where G is the growth

13

rate constant and the 2-D growth exponent is m ~ 2. Correspondingly, the nucleation rate per unit of

14

area varies with time as N0t σ−1, where N0 is the nucleation rate constant and the exponent σ ranges

15

from 0 to 1 for instantaneous and uniform nucleation, respectively.

16 17

Clathration kinetics at the recrystallization front

18

The driving force of the GH growth (crystallization) on ice spheres at the ice-hydrate interface is the

19

local difference between chemical potentials of ice (µi) and hydrate (µw) water frameworks with the

20

former one essentially higher than that of the newly formed hydrate. Based on Eq. (1), a general

21

kinetic equation for the mean ice-to-hydrate conversion rate ωV (the number of ice moles transformed

22

to hydrate on a unit area of ice-hydrate interface per a unit of time) can be formulated as

23

ωV = −

ρ i ri 0 K R ν 1 +ν 2

 1 − y1i 1 − y 2i ν 1 ln + ν 2 ln 1 − y1d 1 − y 2d 

  . 

(12)

24

Here KR is the recrystallization rate constant. The SC- and LC-occupancies y1i, y2i and respective

25

apparent gas fugacities f1i, f2i are ascribed to the clathration reaction front, and they are not

26

necessarily the equilibrium ones, i.e. f1i ≠ f2i in general. For a single-gas clathrate compound its ACS Paragon Plus Environment

11

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 28

1

dissociation pressure pd(T) is an increasing function of temperature T, and the corresponding

2

apparent (equilibrium) gas fugacity f,d(T) and occupancies y1d, y2d are related by analogues of

3

Eqs. (5). The degree of cage equilibration φ1i is introduced as the ratio between the SC- and LC-

4

fugacity deviation with respect to the decomposition fugacity fd:

5

φ1i = ( f 1i − f d ) /( f 2i − f d ) ,

f 1i = f d + φ1i ( f 2i − f d ) .

(13)

6

Thus, for any fixed φ1i-value considered as a tuning parameter, the clathration reaction rate (the

7

driving force) in Eq. (12) depends only on the apparent gas fugacity f2i, i.e. on the gas occupancy y2i

8

in larger cages with the corresponding SC occupancy y1i calculated for f1i. The cage equilibration

9

degree φ1i ranges from 0, at dissociation occupancy y1d and fugacity f1i = fd in SC-s, to 1, at the cage

10

equilibrium conditions when f1i = f2i. One can speculate that φ1i is essentially less than 1 at minimum

11

occupancy level of hydrate stability in SC-s and rather high occupancy in LC-s translated through

12

larger cages from the applied environmental conditions.

13

To explore the kinetic equations (12) and (13), the gas occupancy y2i in larger cages at the reaction

14

front must be specified so as to match the clathration rate with the diffusive gas flux coming to the

15

ice-hydrate interface. This brings forward the necessity to formulate the gas diffusion model.

16 17

Hole-in-cage-wall diffusion sub-model

18

As described by Staykova et al3, a hydrate layer growing around an ice core can be divided by the

19

spherical boundary of radius rc (i.e. by the distance from the ice core centre to an average contact

20

plane) in two sub-layers: a spherical one around the ice core and a layer truncated by the interparticle

21

contacts. Accordingly, the normalized cross-flow area A of the guest-gas diffusive flux on a spherical

22

surface of relative radius R = r/ri0 in the hydrate layer is

23

 R2, Ri < R < Rc ,  A( R) =  2  Z  R c   R 1 − 2 1 − R  , R c < R < R h ,   

 2(1 − s )  , R c = R h 1 − Z  

24

where R i , R c , and R h are the respective analogues of ri, rc, and rh spatially averaged over the

25

hydrate covered area of ice cores and normalized by ri0; Z is the current particle coordination number

ACS Paragon Plus Environment

12

Page 13 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

1

in the sample, and s is the fraction of the open (hydrate/ice) particle surface area exposed to the

2

ambient gas (see also Appendix A of Supporting Information).

3 4 5

At constant diffusion coefficient Dg of the guest molecules moving through water vacancies in LC walls, the general gas transport model (11) can be rewritten in one-dimensional approximation as

ν 2χg ∂y1 [C1 y 2 (1 − y1 ) − C 2 y1 (1 − y 2 )] , = ∂t C1 + C 2

D g ∂  ∂y 2  ∂y 2 ν 1 ∂y1 + = 2 A  . ∂t ν 2 ∂t ri 0 A ∂R  ∂R 

(14)

6

Eqs (14) must be completed by the boundary condition which determines the apparent gas fugacities

7

of the newly formed hydrate so as to counterbalance the GH-growth rates by the gas flux at the ice-

8

hydrate interface, R = R i .

ωV =

9

ρ iν 2 D g ∂y 2 ri 0 c gi

∂R

,

(15)

R = Ri

10

where c gi = ν 1 y1i + ν 2 y 2i is the molar concentration of the guest-gas in the GH at the reaction front.

11

Together with Eqs. (12) and (13), the latter equation (15) predicts the clathration rate and gas

12

occupancies at the ice-core surface.

13

The physics of the initial hydrate layer formation on ice sample surface is not clear in details. The

14

ice surface coating is very difficult to measure with analytical methods, and there is no reliable

15

information on this formation step in terms of cage occupancies. Even for neutron diffraction data

16

the error bars are substantial (Fig S2). Therefore, different deviations of SC- and LC-occupancies,

17

i.e. guest gas fugacities f1h and f2h, from the equilibrium ones corresponding to the applied gas

18

fugacity fa can be envisaged. Thus, as before at the reaction front in Eq. (13), we introduce the

19

equilibration degrees φ1h and φ2h,

20 21

f 1h = f d + φ1h ( f a − f d ) ,

f 2h = f d + φ2h ( f a − f d ) ,

and calculate the corresponding occupancies y1h and y2h, using Eqs. (5).

22

Further outward growth of the hydrate shell occurs due to the outflow of the interstitial excess

23

water molecules from the ice-to-hydrate transformation front. Thus, at the contact with the gaseous

24

environment, we write after Appendix A (Supporting Information) the mass balance relation which

25

governs the outward growth of the particle outer radius R h ,

ACS Paragon Plus Environment

13

Energy & Fuels

Page 14 of 28

2

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

d Rh E Ri d Ri =− , 2 dt s R h dt

1

R h t =0 = 1 ,

(16)

2

and assume that this is a slow equilibrium process proceeding at equilibrium cage filling y1a and y2a,

3

corresponding to the applied gas fugacity fa (see Eqs. (2)).

4 5 6

Eqs. (12)-(16) form a generalized microscopic shrinking-core model, based on the hole-in-cagewall transport mechanism for the GH growth from ice powders during stage II. Finally, the relationship between the clathration degree α and ice-core radius R i , R i = (1 − α / α S )1 / 3 ,

7

(17a)

8

together with the master (mass balance) equation of the ice-to-hydrate conversion on macro-scale

9

level,

10

dα 2 = S i 0 [(1 − α S ) ρ i δ 0 Ω S + α S R i ωV ] , dt

α t =0 = 0 ,

(17b)

11

complete the model of the sI-hydrate formation from ice powders. Here in Eq. (17b) S i 0 = 3 /( ri 0 ρ i )

12

is the initial specific surface area of ice spheres in the sample. Although introduced independently,

13

Eqs. (17) are closely related to the previously developed model2 and, as explained in Appendix C

14

(Supporting Information), can be directly deduced from this model.

15

In view of the general complexity of Eqs. (12)-(17), one has to think carefully about additional

16

observables that can be extracted from in-situ diffraction experiments to constrain the hole-in-cage-

17

wall diffusion model. While the fractions of the formed clathrate phase are reliably determined by

18

neutron measurements, the overall changes in SC/LC filling, being coupled with the thermal

19

displacements, can be derived with considerably lower accuracy, especially at reaction degrees lesser

20

than 10 wt.% (see Experimental). Still, the data remain useful to provide additional constraints for

21

the modeling. Ex-situ Raman or synchrotron X-ray measurements may also give further insights into

22

the composition of the newly-formed gas hydrate and could complement the information obtained

23

from neutron diffraction data.

24 25

Results and Discussion

ACS Paragon Plus Environment

14

Page 15 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

1

Three CO2-hydrate formation runs (see Table 2) starting from deuterated ice particles of 26 µm mean

2

radius at 263, 253, and 230 K and followed by neutron diffraction are considered below. They had

3

been previously simulated and interpreted on the basis of POWDER-4 program as the diffusion-

4

limited clathration process2. Here we continue this analysis with the use of the computer system

5

POWDER-5 which implements the generalized diffusion model given by Eqs. (12)-(17). Further

6

details on this implementation can be found in Appendix D (Supporting Information).

7 8

Diffusion limited ice-to-hydrate conversion

9

A series of preliminary computational experiments by POWDER-5 confirmed the conclusion2 that

10

CO2-hydrate growth on ice spheres was limited by diffusion. The best-fit equilibration factors, mass

11

exchange and diffusion coefficients χg and Dg, initial film thickness δ0, and other model parameters

12

deduced at KR → ∞ are gathered in Table 2. They allow direct estimation of the phenomenological

13

earlier introduced2 permeation coefficients,

14

D′ =

ν 2 ( y 2a − y 2d ) Dg . 2 c gi (1 + ν 0 ) ln( f a / f d )

15

The newly calculated permeation coefficients D' and initial film thickness δ0 for the diffusion-limited

16

CO2-hydrate formation (i.e. at the zero limit of the conversion-type factor F = D ′(1 + ν 0 ) /(ri 0 K R )

17

introduced in our earlier publications2-4) are compared in Table 2 to the previously inferred values

18

(given in parentheses). Close agreement between these principal parameters shows the applicability

19

of the phenomenological POWDER-2 and -4 models2-4 and their consistency with the newly

20

developed generalized theory of the guest-gas diffusion in hydrates.

21

Fig. 1 illustrates the very good match of the neutron diffraction data and performed simulations (STD

22

~0.1-0.3 wt.%). However, the model constraining procedure and sensitivity tests showed that the

23

overall kinetic curves are not noticeably affected by the finer details of the guest-gas migration such

24

as equilibration degrees and LC/SC gas exchange. With this in mind, estimates related to the SC and

25

LC filling variation during the clathration reaction (Fig. 2) have been deduced from neutron

26

measurements in the case of CO2-hydrate formation2-4. At the initial, coating phase of the formation2,

2

ACS Paragon Plus Environment

2

15

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 28

1

below 5-7 wt.% of GH (after 12-18 min), we have observed increased uncertainties and larger scatter

2

of data points (see Fig. S2 for plots with error bars). A similar situation is also noticeable for the

3

whole formation run at 230 K where the conversion degree and reaction rate are low. The

4

occupancies of all investigated cases display clear trends; occupancies of LC-s rapidly settle at ~0.97

5

and SC-s show a continuous, slow readjustment during the whole reaction length. Cage occupancies

6

have allowed a first-guess constrain of all the major hole-in-cage-wall diffusion parameters,

7

including the equilibration factors and guest-gas mass exchange coefficients; the resulting best-fit

8

kinetic curves are also plotted in Fig. 2.

9

It should be noted that, in spite of the refinement constraints, the experimentally obtained absolute

10

cage occupancies show a minor overestimation with respect to the simulated curves about ~ 2 % for

11

the LC-s and ~ 5-9 % for SC-s. This is somewhat larger than the conditional 1σ error from the

12

constant Rietveld refinement (Fig. S2), yet still remains within the marginal error (i.e. including

13

uncertainties from the constrained displacement parameters). The reason for this discrepancy in

14

principle might be related either to the imperfect constraints of the data (mainly, atomic displacement

15

parameters) or to poorly defined Langmuir constants used in the model. Sensitivity tests on the first

16

possibility showed that small variation of constraints in the Rietveld refinement shifts the absolute

17

values of cage occupancies by a few % without noticeable distortion of relative changes and general

18

trends. At the same time, similar trails on the Langmuir constants allowed very little flexibility,

19

leading to unphysical diffusion rates. Hence, to reliably adjust the slopes of the simulated curves and

20

come to consistency between the data and model predictions, we simultaneously shifted the deduced

21

SC and LC measurements by the best-fit constant values (see Fig. S2).

22

For the diffusion-limited CO2-hydrate formation from ice powder, the SC and LC occupancies at

23

the reaction front are very close to the dissociation levels; the equilibration degree φ1i does not affect

24

the ice-to-hydrate conversion and has been set as φ1i = 0. Rather high guest-gas mass exchange

25

coefficients, on the order of 10-35 h−1, are estimated and reveal a fast relaxation of the depleted SC

26

occupancies to equilibrate with the local LC filling. The overall diffusion flux passing through LC-s

27

is high enough to compensate for excess or lack of guest molecules and quickly equilibrates the

ACS Paragon Plus Environment

16

Page 17 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

1

initial hydrate patches with the gaseous environment. Thus, we have had to assume φ2h = 1. Although

2

systematically small, only the initial SC equilibration degree φ1h seems to play (see Table 2) a certain

3

role as a controlling model parameter.

4 5

Reaction-controlled ice-to-hydrate conversion

6

The lowest possible values of the clathration rate constant KR and maximum bounds for the diffusion

7

coefficients Dg have been also estimated in sensitivity tests by attempts to fit the available data of the

8

three CO2-hydrate formation Runs 1-3 at highest non-zero conversion type factors F (see above)

9

which allowed the same accuracy level. The limiting best-fit model parameters have been obtained at

10

F ~ 0.06-0.07 and are summarized in Table S2; the previously deduced values2 are given in

11

parentheses. The measurements and model predictions are compared in Figs. S3 and S4. Fig. S3

12

illustrates still a good agreement between the performed simulations and the experimental values of

13

the reaction degree of a similar quality level as before in Fig. 1 for the diffusion-limited scenario.

14

The reduced reaction rate constants enhance somewhat the simulated variations of averaged SC-

15

occupancies (compare Fig. 2 and Fig. S4) but do not noticeably influence the deviations from

16

experimental LC and SC neutron diffraction measurements. Again, we have to emphasize that the

17

experimental occupancy data are not sufficiently accurate to reveal further details of the ice-to-

18

hydrate conversion especially at low reaction degree, below ~ 5-7 wt.% ( starting 12-18 min of the

19

reaction), in particular for low-temperature Run 3. Nevertheless, the performed computational

20

exercise shows that our estimates of the diffusion coefficients presented in Table 2 does not change

21

by more than a factor of 2 for the runs at 253 and 263 K and by more than a factor of 3 for the low

22

temperature run at 230 K, while all other model parameters remain practically same.

23

Although there are some uncertainties in the interpretation of the experimental data, our findings

24

suggest that the essential parameters of interest, in particular the diffusion constants, can be

25

determined from the shrinking-core modeling. Our estimates are clearly better than an order-of-

26

magnitude approximation. The work reported here is only concerned with CO2 diffusion in a hydrate

27

host lattice; yet, the results can still be compared to our earlier analysis of the CH4 diffusion, since

28

our earlier phenomenological approaches are in good agreement with the more sophisticated and 17 ACS Paragon Plus Environment

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 28

1

physical “hole-in-cage-wall” model. It is noteworthy in this context that the former permeation

2

model was established largely considering the methane case. Thus, the ~ 3-times higher bulk

3

diffusion constants of CO2, as compared to CH4 2, would likely persist also for the CH4-case analysis

4

with our new physical model; the small influence of cage filling on the diffusion constants16 will not

5

alter this picture, despite the fact that the CO2-hydrate shows a 5% smaller total cage filling as

6

compared to CH4-hydrate at identical p-T conditions. The inferred diffusion constants could, thus, be

7

inserted into the model for gas transport phenomena in GH, e.g. in gas-exchange reaction related to

8

the replacement of CH4 for CO2, a process which would allow the concomitant sequestration of CO2

9

and recovery of CH4 from natural gas hydrates e.g.36-44.

10 11

Mid- and post-formation equilibration

12

A direct consequence of the diffusive transport of guest molecules across the GH layer is a

13

continuous readjustment of SC- and LC- occupancies. The highest difference in cage occupancies

14

between the hydrate surface (assumed to equilibrate more or less instantaneously in our experimental

15

time-resolution) and the reaction front is observed in case of the diffusion-limited scenario. The

16

compositional gradient develops after formation of the initial film of thickness δ0 (see Tables 2 and

17

S1) and remains until the equilibrium with the gas phase in whole volume of GH is reached. What

18

seems to be insufficiently appreciated in the past is the fact that the equilibration length extends

19

beyond the point of the full transformation (understood as a full conversion of ice into gas hydrate).

20

This fact is of particular importance for estimation of cage occupancies in not fully converted or not

21

fully equilibrated samples that will be smaller than predicted equilibrium values.

22

To demonstrate the duration of the readjustment process, we use the results from 263 K (Table 2,

23

Run 1) to simulate the time necessary for the full equilibration of GH grown (Fig. 3) in case of

24

26 µm-mean-radius particles. The complete transformation to CO2-hydrates takes about 470 h (i.e.

25

~19 days), after that practically uniform occupancy distributions, within 0.1‰ resolution of the

26

model, are attained in 50 and 110 h (in ~2-5 days) for LC-s and SC-s, respectively. For larger

27

particles the equilibration period increases proportionally to the time necessary for the full

28

conversion (Fig. 4). For ice spheres with the mean radius of 100 µm the formation takes nearly 290 ACS Paragon Plus Environment

18

Page 19 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

1

days for completion and additional 20 to 40 days for the equilibration of large and small cages. Thus,

2

the presence of large ice grains in the starting material will easily lead to an incomplete conversion

3

into GH, which is frequently observed2-4; it should be mentioned that ice grain polydispersity on its

4

own also accounts for some relic ice as a consequence of switching larger grains out of the reaction -

5

a consequence of the outward growth of GH on ice grains2, 4. Care must also be taken concerning the

6

homogeneity of GH samples prepared with the ice method. Extended reaction times are needed for a

7

full equilibration and homogenization of the cage fillings when starting from ice. Discrepancies

8

between cage fillings determined by X-ray diffraction, Raman spectroscopy or NMR on laboratory

9

specimens (obtained via the ice method) and natural samples (with presumably longer equilibration

10

times) are to be expected. Likewise, discrepancies between experimental values and theoretical

11

predictions of cage fillings may simply be due to insufficient equilibration.

12 13

Scale analysis of the phenomenological models2-4 and their consistency with the newly developed generalized “hole-in-cage-wall” diffusion model allows to introduce the dimensionless time

14

Fo =

15

6tν 2 ( y 2 a − y 2 d ) D g c g i ri 0

2

=

6tD ′ (1 + ν 0 )r i 0 2

2

ln

fa , fd

16 17

which in case of diffusion limited ice-to-hydrate conversion at relatively high SSA-equivalent

18

particle radii ri0 > 50-70 µm (i.e. mean radii > 20-30 µm) controls the hydrate formation

19

process and the post-formation equilibration. Using the deduced model parameters from Table 2, it

20

becomes possible to present all the principal simulated normalized kinetic curves (e.g. from Figs. 1

21

and 2) for Runs 1-3 at 263, 253, and 230 K versus Fo for different initial particle radii in a unified

22

practically identical form (Fig. 5); normalization of the cage filling was achieved by

23

putting y j = ( y j − y j d ) /( y j a − y j d ) , j = 1, 2. In particular, it can be concluded that the ice-to-

24

hydrate conversion time is about Fo ~ 1.08-1.13, while the equilibration is essentially finished at

25

Fo ~ 1.17-1.30. These estimates slightly vary due to differences in the deduced equilibration degrees

26

φ1h ~ 0-0.1. ACS Paragon Plus Environment

19

Energy & Fuels

Page 20 of 28

1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

2

CONCLUSIONS

3

1. A generalized shrinking core model for sI-hydrate formation from ice powders is developed on

4

the basis of hole-in-cage-wall gas diffusion through the hydrate layer to reaction front in mono-size

5

approximation valid up to 40-50% of ice-to-hydrate conversion.

6

2. The interactive computer system POWDER-5 is implemented to simulate the ice-to-hydrate

7

conversion in the framework of the hole-in-cage-wall diffusion concepts and to interpret the hydrate

8

formation kinetics data.

9

3. Model sensitivity analysis showed that, apart from the overall reaction degree curves, additional

10

data on spatial and temporal LC- and SC-filling variations in the growing hydrate layers would be

11

needed to reliably reveal the transient cage-occupancy readjustment and details of the inter-structural

12

gas exchange mechanisms in hydrates. Such data may be obtained from Raman spectroscopy45, 46.

13

4. The evolution of the volume-averaged LC- and SC-occupancies may be considered as one of the

14

most informative data to deeper understand and constrain the diffusion (gas mass transfer) processes

15

in gas hydrates. However, the acquisition of sufficiently accurate (neutron diffraction) data,

16

especially at reaction degrees below 5-7 wt.%-level is challenging from the instrumental point of

17

view. Weakly scattering guest molecules like CH4 or H2S will hardly be accessible by this method.

18

5. The principal model parameters inferred from the kinetic data in case of the CO2-hydrate

19

formation, such as initial hydrate-layer thickness and permeation coefficients, remain in close

20

agreement with those deduced earlier on the basis of the simplified phenomenological approach

21

employed in the POWDER-4 software.

22

6. While the modeling of the initial stages of the clathration reaction is sensitive to experimental

23

variability and bound with uncertainties of the data analysis, the later permeation-controlled stage is

24

less prone to such problems and delivers reliable kinetic parameters.

25

7. The POWDER-5 model offers a unique insight into the duration of the equilibration process of

26

the cage fillings; the time-scale of this process exceeds the time of the total ice-to-hydrate

27

conversion.

ACS Paragon Plus Environment

20

Page 21 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

1

8. The gas diffusion constants deduced from GH formation experiments are considered to be

2

relevant also to natural and technologically motivated gas exchange processes as the underlying hole-

3

in-cage-wall model is expected to be closely similar for a specific guest in a given host framework;

4

the results presented here are valid for GH crystallizing in structure type I

5

9. While it is very likely that the methane case is analogous to the CO2 “hole-in-cage-wall”

6

mechanism, the situation for larger non-polar (e.g. Xe) and for polar guest molecules in general (e.g.

7

H2S) remains to be elucidated and merits further efforts. Any H-bonding between guest and host may

8

significantly alter the concentration of water vacancies and thus affect guest permeabilities;

9

molecular dynamics simulations are likely to provide useful insights into such guest-specific

10

differences.

11 12

ACKNOWLEDGEMENTS

13

The authors thank the Institut Laue-Langevin (ILL) in Grenoble/ France for beam time and support.

14

Financial support was granted by BMBF in the framework of its SUGAR-II program (grant

15

03G0819B, TP B2-3). The technical help of Ulf Kahmann and Heiner Bartels (both Göttingen) is

16

gratefully acknowledged.

17

ACS Paragon Plus Environment

21

Energy & Fuels

Page 22 of 28

1 230K 253K 263K

GH wt. fraction [α]

0.4

0.3

0.2

0.1

0.0 0

5

10

15

20

25

30

Time [h]

2 3 4 5

Figure 1 Kinetics of CO2-hydrate formation from deuterated ice powder at 263, 253, and 230 K in Runs 1, 2, and 3 (circles) of neutron diffraction experiments and the best-fit kinetic curves (solid lines) simulated by POWDER-5 system in the diffusion limited scenario.

6

230K 253K 263K

1.04

1.02

0.55

Cage occupancy (SC)

Cage occupancy (LC)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1.00

0.98

0.96

0.50

0.45

0.40

0.94 0

5

10

15

20

25

30

0

5

10

Time [h]

15

20

25

30

Time [h]

7 8 9 10 11 12

Figure 2 Variation of occupancies versus time for LC-s and SC-s: CO2-hydrate formation from deuterated ice powder at 263, 253, and 230 K and the best-fit kinetic curves (solid lines) simulated by POWDER-5 system in the diffusion-limited scenario. LC occupancy data > 1.0 are unphysical and have been dimmed. The uncertainty of each single entry is up to 0.05 for the initial part of the reaction and becomes smaller as the conversion degree increases (see Fig. S2 for enlarged plots with error bars).

13 14 15 16 ACS Paragon Plus Environment

22

Page 23 of 28

Energy & Fuels

1 0.972

2

0.971 0.970

3

5 6

Cage occupancy

4

(2)

LC-s

0.969

(3)

(1)

0.968 0.967 0.966

0.52

(2)

SC-s

7 8

(3)

(1)

0.48

0

100

200

300

400

500

600

Time [h]

9 10 11 12 13 14

Figure 3 Simulated equilibration path of cage occupancies in GH grown from ice powders of 26-µm-mean radius at 263 K and 1 MPa: (1) during the GH formation, (2) after the complete transformation but before the equilibration; (3) The full equilibration (within 0.1‰); The readjustment is predicted to be somewhat faster for LC-s than for SC-s.

15 16 17 18

263 K 1 MPa

Full conversion & equilibraiton [days]

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

300

250

200

150

100

SC LC Full conversion

50

20

40

60

80

100

Particle radius [µm]

19 20 21

Figure 4 Calculated time of the full conversion and LC/SC equilibration as a function of the mean particle radius at constant temperature of 263 K and pressure of 1 MPa.

22

ACS Paragon Plus Environment

23

Energy & Fuels

Page 24 of 28

1 2

1.0

0.8

Cage occupancy

0.8

0.6

0.4

0.6

0.4

0.2

0.2

0.0 0.0

1.0

230 K 253 K 263 K

Ri

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

0.2

0.4

0.6

0.8

1.0

1.2

1.4

0.0 0.0

230 K 253 K 263 K 0.2

LC LC LC 0.4

SC SC SC

0.6

0.8

1.0

1.2

1.4

Fo

Fo

3 4 5 6 7

Figure 5 Characteristics of CO2-hydrate formation from deuterated-ice particles ( = 26 µm, r i 0 = 42.6 µm, ri0 = 69.9 µm) and post-conversion equilibration (Fo > 1.1) for Runs 1-3 at 263, 253, and 230 K (see Table 2) simulated by POWDER-5 system in diffusion limited scenario: (l.h.s) ice-core radius Ri, and (r.h.s) normalized volume-averaged SC and LC occupancies.

8 9 10 11 12 13 14 15 16 17

ACS Paragon Plus Environment

24

Page 25 of 28

Energy & Fuels

1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Characteristics

sI

No. of H2O molecules in the unit cell

46 11.8512.05 43.8 2 5.1 1/23 6 5.8 3/23

Unit cell edge (lattice constant), Å Water density ρw, kmol/m3 Number of smaller cages (SC) SC free diameter, Å SC number per one water molecule ν1 Number of larger cages (LC) LC free diameter, Å LC number per one water molecule ν2

2 3

Table 1 The unit cell characteristics of the sI clathrate water frame.

4 5 Conditions of experiments T K 1

263

2

253

3

230

p(f) MPa 1.0 (0.93) 1.0 (0.91) 0.3 (0.29)

pd(fd) MPa 0.75 (0.71) 0.51 (0.49) 0.186 (0.183)

Deduced kinetic parameters h-1

χg

Dg m2/h

D' m2/h

0

10

2.45⋅10-10

4.26(4.3)⋅10-12

1.21⋅107

0.1

35

1.62⋅10-10

2.07(2.1)⋅10-12

2.5⋅107

0

25

2.98⋅10-11

2.71(2.7)⋅10-13

C1 MPa-1

C2 MPa-1

m

kG m/hm/2

σ

kN 1/hσm2

φ1h

1.24

36.1

2

3.5⋅10-3

0.5

1.49⋅107

1.76

61.3

2

11⋅10-3

0.6

4.5

250

1.6

9.5⋅10-4

0.4

δ0 µm 6.26 (6.7) 4.14 (4.3) 1.32 (1.33)

6 7 8 9

Table 2 Diffusion-limited scenario. Conditions of experiments and kinetic parameters of CO2-hydrate formation from ice powders /Simulations by the POWDER-5 system. Previously inferred2-4 parameters are

given in parentheses.

10 11 12 13 14 15 16 17 18 19 20 ACS Paragon Plus Environment

25

Energy & Fuels

1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54

1.

2.

3.

4. 5. 6. 7. 8.

9.

10. 11. 12.

13. 14. 15.

16.

17.

18. 19.

20. 21.

Falenty, A.; Genov, G.; Hansen, T. C.; Kuhs, W. F.; Salamatin, A. N., Kinetics of CO2 Hydrate Formation from Water Frost at Low Temperatures: Experimental Results and Theoretical Model. Journal of Physical Chemistry C 2011, 115 (10), 4022-4032. Falenty, A.; Salamatin, A. N.; Kuhs, W. F., Kinetics of CO2-Hydrate Formation from Ice Powders: Data Summary and Modeling Extended to Low Temperatures. Journal of Physical Chemistry C 2013, 117 (16), 8443-8457. Staykova, D. K.; Kuhs, W. F.; Salamatin, A. N.; Hansen, T., Formation of porous gas hydrates from ice powders: Diffraction experiments and multistage model. Journal of Physical Chemistry B 2003, 107 (37), 10299-10311. Kuhs, W. F.; Staykova, D. K.; Salamatin, A. N., Formation of methane hydrate from polydisperse ice powders. Journal of Physical Chemistry B 2006, 110 (26), 13283-13295. Ishida, Y.; Sakemoto, R.; Ohmura, R., Crystal growth of clathrate hydrate in gas/liquid/liquid system: variations in crystal-growth behavior. Chemistry 2011, 17 (34), 9471-7. Sun, C. Y.; Peng, B. Z.; Dandekar, A.; Ma, Q. L.; Chen, G. J., Studies on hydrate film growth. Annual Reports Section "C" (Physical Chemistry) 2010, 106, 77-100. Watanabe, S.; Saito, K.; Ohmura, R., Crystal Growth of Clathrate Hydrate in Liquid Water Saturated with a Simulated Natural Gas. Crystal Growth & Design 2011, 11 (7), 3235-3242. Li, S. L.; Sun, C. Y.; Liu, B.; Feng, X. J.; Li, F. G.; Chen, L. T.; Chen, G. J., Initial thickness measurements and insights into crystal growth of methane hydrate film. Aiche J. 2013, 59 (6), 2145-2154. Genov, G.; Kuhs, W. F.; Staykova, D. K.; Goreshnik, E.; Salamatin, A. N., Experimental studies on the formation of porous gas hydrates. American Mineralogist 2004, 89 (8-9), 1228-1239. Vlasov, V. A., Phenomenological diffusion theory of formation of gas hydrate from ice powder. Theor. Found. Chem. Eng. 2012, 46 (6), 576-582. Vlasov, V., Diffusion model of gas hydrate formation from ice. Heat Mass Transfer 2015, 17. Liu, W.; Li, Q.; Song, Y.; Zhang, L.; Yang, M.; Wang, L.; Chen, Y., Diffusion Theory of Formation of Gas Hydrate from Ice Powder without Melting. Energy Procedia 2014, 61, 513522. Falenty, A.; Hansen, T. C.; Kuhs, W. F., Formation and properties of Ice XVI obtained by emptying a clathrate hydrate sII. Nature 2014, 516 (232), 231–233. Liang, S.; Kusalik, P. G., The Mobility of Water Molecules through Gas Hydrates. Journal of the American Chemical Society 2011, 133 (6), 1870-1876. Buch, V.; Devlin, J. P.; Monreal, I. A.; Jagoda-Cwiklik, B.; Uras-Aytemiz, N.; Cwiklik, L., Clathrate hydrates with hydrogen-bonding guests. Physical Chemistry Chemical Physics 2009, 11 (44), 10245-10265. Peters, B.; Zimmermann, N. E. R.; Beckham, G. T.; Tester, J. W.; Trout, B. L., Path Sampling Calculation of Methane Diffusivity in Natural Gas Hydrates from a Water-Vacancy Assisted Mechanism. Journal of the American Chemical Society 2008, 130 (51), 1734217350. Hansen, T. C.; Henry, P. F.; Fischer, H. E.; Torregrossa, J.; Convert, P., The D20 instrument at the ILL: a versatile high-intensity two-axis neutron diffractometer. Meas. Sci. Technol. 2008, 19 (3). Larson, A. C.; Dreele, R. B. General Structure Analysis System (GSAS); Los Alamos National Laboratory: 2000; pp 86-748. Hansen, T. C.; Falenty, A.; Murshed, M. M.; Kuhs, W. F. Lattice constants and expansivities of gas hydrates from 10 K up to the stability limit, In: Proceedings of the 8th International Conference on Gas Hydrates, Beijing, China, July 28- August 1, 2014; pp 1-7. Klapproth, A.; Goreshnik, E.; Staykova, D.; Klein, H.; Kuhs, W. F., Structural studies of gas hydrates. Canadian Journal of Physics 2003, 81 (1-2), 503-518. Hartmann, C. D.; Hemes, S.; Falenty, A.; Kuhs, W. F. The structure and cage filling of gas hydrates as established by synchrotron powder diffraction data, In: Proceedings of the 7th 26 ACS Paragon Plus Environment

Page 26 of 28

Page 27 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54

Energy & Fuels

22. 23. 24. 25.

26.

27.

28. 29.

30. 31. 32.

33. 34. 35. 36.

37.

38. 39. 40.

41. 42.

43.

International Conference on Gas Hydrates, Edinburgh, Scotland, United Kingdom, July 1721, 2011. van der Waals, J. H.; Platteeuw, J. C., Clathrate Solutions. Adv. Chem. Phys. 1959, 2, 1-57. Sloan, E. D.; Koh, C. A., Clathrate hydrates of natural gases. 3rd ed.; CRC Press, Taylor & Francis Group: Boca Raton, FL, 2008. Chazallon, B.; Kuhs, W. F., In situ structural properties of N2, O2, and air-clathrates by neutron diffraction. Journal of Chemical Physics 2002, 117 (1), 308-320. Murshed, M. M.; Schmidt, B. C.; Kuhs, W. F., Kinetics of Methane-Ethane Gas Replacement in Clathrate-Hydrates Studied by Time-Resolved Neutron Diffraction and Raman Spectroscopy. Journal of Physical Chemistry A 2010, 114 (1), 247-255. Salamatin, A. N.; Hondoh, T.; Uchida, T.; Lipenkov, V. Y., Post-nucleation conversion of an air bubble to clathrate air-hydrate crystal in ice. Journal of Crystal Growth 1998, 193 (1-2), 197-218. Demurov, A.; Radhakrishnan, R.; Trout, B. L., Computations of diffusivities in ice and CO2 clathrate hydrates via molecular dynamics and Monte Carlo simulations. Journal of Chemical Physics 2002, 116 (2), 702-709. Geil, B.; Kirschgen, T. M.; Fujara, F., Mechanism of proton transport in hexagonal ice. Phys. Rev. B 2005, 72 (1), 014304. Kirschgen, T. M.; Zeidler, M. D.; Geil, B.; Fujara, F., A deuteron NMR study of the tetrahydrofuran clathrate hydrate - Part II: Coupling of rotational and translational dynamics of water. Physical Chemistry Chemical Physics 2003, 5 (23), 5247-5252. Katchalsky, A.; Curran, P. F., Nonequilibrium Thermodynamics in Biophysics. Harvard University Press: Cambridge, Massachusetts, USA, 1965; p 258. Lasaga, A. C., Kinetic Theory in the Earth Sciences. Princeton University Press: Princeton, New Jersey, USA, 1998. Hirohama, S.; Shimoyama, Y.; Wakabayashi, A.; Tatsuta, S.; Nishida, N., Conversion of CH4-hydrate to CO2-hydrate in liquid CO2. Journal of Chemical Engineering of Japan 1996, 29 (6), 1014-1020. Arzt, E., The influnce of an increasing particle coordination on the densification of spherical powders. Acta Metallurgica 1982, 30 (10), 1883-1890. Salamatin, A. N.; Kuhs, W. F. Formation of porous gas hydrates, In: Proceedings of the Fourth International Conference on Gas Hydrate, Yokohama, 2002; pp 766-770. Humphreys, F. J.; Hatherley, M., Recrystallization and Related Annealing Phenomena. 1st ed.; Pergamon Press: Oxford, UK, 2002. Lee, H.; Seo, Y.; Seo, Y. T.; Moudrakovski, I. L.; Ripmeester, J. A., Recovering methane from solid methane hydrate with carbon dioxide. Angew. Chem.-Int. Edit. 2003, 42 (41), 5048-5051. Yoon, J. H.; Kawamura, T.; Yamamoto, Y.; Komai, T., Transformation of methane hydrate to carbon dioxide hydrate: In situ Raman spectroscopic observations. Journal of Physical Chemistry A 2004, 108 (23), 5057-5059. Ota, M.; Abe, Y.; Watanabe, M.; Smith, R. L.; Inomata, H., Methane recovery from methane hydrate using pressurized CO2. Fluid Phase Equilib. 2005, 228, 553-559. Ota, M.; Morohashi, K.; Abe, Y.; Watanabe, M.; Smith, R. L.; Inomata, H., Replacement of CH4 in the hydrate by use of liquid CO2. Energy Conv. Manag. 2005, 46 (11-12), 1680-1691. Schicks, J. M.; Luzi, M.; Beeskow-Strauch, B., The Conversion Process of Hydrocarbon Hydrates into CO2 Hydrates and Vice Versa: Thermodynamic Considerations. Journal of Physical Chemistry A 2011, 115 (46), 13324-13331. Deusner, C.; Bigalke, N.; Kossel, E.; Haeckel, M., Methane Production from Gas Hydrate Deposits through Injection of Supercritical CO2. Energies 2012, 5 (7), 2112-2140. Yuan, Q.; Sun, C. Y.; Yang, X.; Ma, P. C.; Ma, Z. W.; Liu, B.; Ma, Q. L.; Yang, L. Y.; Chen, G. J., Recovery of methane from hydrate reservoir with gaseous carbon dioxide using a threedimensional middle-size reactor. Energy 2012, 40 (1), 47-58. Falenty, A.; Murshed, M. M.; Salamatin, A. N.; Kuhs, W. F. Gas Replacement in Clathrate Hydrates during CO2 Injection - Kinetics and Micro-Structural Mechanism, In: Tenth ISOPE ACS Paragon Plus Environment

27

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1 2 3 4 5 6 7 8 9

44. 45. 46.

Page 28 of 28

Ocean Mining and Gas Hydrates Symposium, The International Society of Offshore and Polar Engineers (ISOPE): Szczecin, Poland, 22-26, September, 2013; pp 109-115. Goel, N., In situ methane hydrate dissociation with carbon dioxide sequestration: Current knowledge and issues. J. Pet. Sci. Eng. 2006, 51 (3–4), 169-184. Qin, J. F.; Kuhs, W. F., Quantitative Analysis of Gas Hydrates Using Raman Spectroscopy. Aiche J. 2013, 59 (6), 2155-2167. Qin, J.; Kuhs, W. F., Calibration of Raman Quantification Factors of Guest Molecules in Gas Hydrates and Their Application to Gas Exchange Processes Involving N2. Journal of Chemical & Engineering Data 2014.

10

ACS Paragon Plus Environment

28