You are on page 1of 12

Phenomenological Modeling

of Hydrate Formation and


Dissociation
Maria Carolina Gonzalez Chacin,
Richard G. Hughes,* Faruk Civan,
and Charles E. Taylor
1. INTRODUCTION
Since the pioneering work of Hammerschmidt [1], the hydrate formation and
dissociation phenomena have been the subject of numerous studies (Sloan, [2],
and Makogan, [4]). Englezos, et al. [5,6]) proposed a model for hydrate
ormation based on experimental data they collected on the formation of
hydrates of methane, ethane, and their mixtures at constant pressure or
temperature. The turbidity time was identified visually as a transition from a
clear to translucent solution. An induction period was defined, beginning with
the initiation of the experiment until the attainment of the turbidity point. The
dominant process during the induction time was assumed to be physical
absorption of the gas in the liquid. A growth period was then noted which
started at the turbidity point and continued until theexperiment concluded. The
experimental data collected during this period was also considered in modeling.
The overall driving force for the crystallization process was assumed to be the
difference between the dissolved gas fugacity and the three-phase equilibrium
fugacity at the experimental conditions.
Carrying out a parametric analysis with the Englezos et al. model [5,6] and
experimental data, Skovborg and Rasmussen [7] concluded that the agreement
between some of the experimental data and the model was questionable. They
observed that the model performed well on the early-time data, but that there
were discrepancies with the late-time data. The model indicated increasing rate
of reaction with time, while the experimental data did not reflect such behavior.
Examining the other available experimental information (Bishnoi et al., [7,8]),
Skovborg and Rasmussen [7] noticed that the reaction rate was constant or, in
some cases had a declining tendency. The transfer of gas from the gas phase
to or through the liquid phase was considered to be the determining step. The
discrepancies between experimental data and the Englezos, et al. Model is
because the total reaction rate was assumed proportional to the second
moment of the particle size distribution, which in effect determines the total
surface area of the particles. This area increases as the particles grow and
therefore the reaction rate increases with time. Also, Englezos, et al. neglected
the interactions between hydrate particles in the population mass balance,
assuming that the particles would always be small in comparison to the medium
containing them. However, the number of particles and their size will grow and
eventually the particles will interact with each other. These interactions may turn
into a coagulation process when the container space restricts their growth.
Thus, at longer times, hydrate particles will be a part of a continuous medium
forming a hydrate bulk phase. Hence, the total surface area per volume of the
particles will decrease and consequently the probability of interactions between
the reactants will also decline. The low gas permeability of the hydrate and the
low diffusion coefficient of methane in water are the key factors limiting the
mass transfer and the hydrate reaction.
Natarajan et al. [8] presented some induction data measurements for formation
of methane, ethane, and carbon dioxide hydrates. These measurements were
obtained in the same equipment using the same experimental procedure as in
Englezos, et al. [5]. A meta-stable region for hydrate nucleation was considered.
They used the crystallization theory to formulate a model for the hydrate
induction time required for formation of stable hydrate nuclei. Makogon [9] and
Englezos, et al. [5,6] pointed out the similarities between the crystallization and
hydrate formation processes. Thus, the hydrate formation process was divided
into two periods: “nucleation” and “growth”. During the nucleation phase,
hydrate nuclei are generated from a super-saturated aqueous solution. The
nucleation process ends by the appearance of critical-sized stable hydrate
nuclei.
Figure 1. Typical hydrate nucleation and growth trends (after Natarajan et al., [8]).

Natarajan et al. [8] measured the moles of gas consumed with time during a typical
hydrate formation experiment. They observed trends schematically described in Figure
1. The growth period was assumed to start after the “turbidity time”. A slight dip in the
reactor pressure occurred at the turbidity point due to the sudden loss of super-saturation
in the liquid. A sharp discontinuity was also observed in the curve of moles of gas
consumed vs. time, with a slight rise in the solution temperature at the turbidity point,
due to the heat release during the stable nuclei formation. Natarajan et al. [8] obtained
their experimental data at isothermal and isobaric conditions only during the induction
time. The measured induction times for methane showed an exponential tendency with
the fugacity-based driving force.
An examination of the various experimental data including the NETLgenerated data
processed in this study reveal three basic insights about the hydrate formation process:

The hydrate formation process occurs in a three-step sequence:


1. Transport of gas from the gas phase to the aqueous liquid phase.
2. Diffusion of gas from the aqueous phase and through the liquid film to the hydrate-
liquid interface.
3. Reaction of gas with the aqueous phase to form hydrates at the interface.

There is a critical radius of the hydrate nuclei. Nuclei having a radius greater than the
critical radius are stable and continue to grow, while the nuclei with radius smaller than
the critical radius re-dissolve in the liquid medium.
The rate of the phase transformation is directly related to the three-phase equilibrium
pressure. The deviations from the three-phase equilibrium will generate conditions
necessary for formation and thus the reaction takes place.

Here, a simple model for analysis of experimental data is developed in view of these
insights. Both equilibrium and non-equilibrium models are presented and the data of
Makogan and Holditch [9,10] and the present in-house experimental data are analyzed
using these models. Our studies indicate that hydrate formation is best described using
the non-equilibrium model while the equilibrium model best describes the hydrate
dissociation process. The present model is not intended to be predictive; rather it
provides a means of analyzing experimental data in order to gain insights into the
mechanism of the hydrate formation and dissociation processes.

2. PROCESS DESCRIPTION
The proposed model considers a closed reactor, where methane and water are brought in
contact at low temperature and high pressure, to form a hydrate. The formation process
is divided into three time periods (See Figure 1): dissolution, induction, and growth
periods. During the first and second periods, the gas and liquid phases are present in the
chamber. During the dissolution period, the governing process is the mass transfer from
the gas to the liquid. This period ends when the gas saturates the liquid. Hence, the
concentration of the gas in the water increases, and super-saturated regions appear,
leading to a clustering of water molecules around gas molecules. The induction period
ends when clusters attain a critical size. Natarajan et al. [8] modeled the induction and
dissolution periods with the same approach although the dominant processes involved
during the dissolution and induction periods are different. The third time period is
defined as the growth period. A three- phase system, consisting of the liquid, gas, and
hydrate, is present. It is assumed that an excess amount of gas is available in the
chamber during the whole process, and the gas volume reduction is negligible, given
that the change of volume of water at different pressures is negligible and the volume
occupied by the solution reacted and the hydrate formed are similar. Figure 2 shows a
schematic of the processes involved during this period.
Although the liquid can achieve uniform super-saturation before the appearance of the
hydrate nuclei (Englezos et al., [5]), experimental evidence indicates that the nucleation
and growth occur predominantly at the gas-liquid interface, where the super-saturation
is higher than anywhere else in the liquid. The appearance of a hydrate layer between
the gas and liquid phases depends on the mixing conditions of the system and the
amount of un-reacted water present in the system.

2.1. Equilibrium Model


The Clapeyron equation describes the equilibrium vapor pressure dependence on
temperature. It can be applied to any kind of phase equilibrium including

Figure 2. Mass transfer phenomenon during growth period at constant temperature and pressure.

solid-vapor (Elliot and Lira, [11]) as long as the system is univariant, as is the case for
simple hydrates (van der Waals and Platteuw, [12]). This relationship can be expressed
as:

where P and T are the prevailing pressure and temperature, is the heat of
formation/dissociation, Z is the gas deviation factor and R is the universal gas constant.
Thus, plotting In P vs. 1/T will generate a straight line. Sloan [2]) stated that such linear
plots might either indicate a constant of hydrate formation when the
compressibility factor, and the stoichiometric ratio of water and the guest molecule can
be considered constants, or the non-linear behaviors of these factors cancel each other.
The slope for the formation and dissociation should be the same for equilibrium
processes. The statistical thermodynamic software package Hydoff (Sloan, [3]) was
used to predict the equilibrium hydrate pressure for a given temperature, and these
results were compared with the experimental pressures. This program assumes
equilibrium between the three phases, and has a stated error of 5% when compared to
experimental data.

2.2. Non-equilibrium Model


The in-house data calculations (see results for equilibrium approach calculations)
indicated that the hydrate formation process did not follow an equilibrium path. Also
Sloan [2] pointed out that the Clapeyron equation presents

Figure 3. System Configuration for: (a) Dissolution and Induction Periods and (b) Hydrate Growth Period.

discrepancies with the experimental values because hydrate formation is


nonstoichiometric. Such discrepancies are not important at equilibrium conditions
because the non-stoichiometry remains approximately constant over small temperature
ranges. The temperature ranges involved in the experimental data analyzed here are
large. Therefore, a different approach is necessary to analyze such data.
We propose a non-equilibrium approach to analyze the data that is based on the process
description presented earlier. The conceptual system configurations during the
dissolution and induction periods, and the hydrate growth period are shown in Figure 3.
The gas and liquid phases initially occupy certain prescribed regions of the reactor
(Figure 3a). For the induction and growth periods, all hydrate particles are assumed to
migrate to the gas-liquid interface rapidly and grow there (Figure 3b).
2.2.1. Gas Phase Equations
Neglecting the water vapor pressure, the cumulative gas consumption can be expressed
as the difference between the initial and instantaneous moles of methane in the gas
phase:

During the dissolution time period, this gas consumption is assumed to be going into the
liquid phase. The discussion below will present equations defining how the
consumption is handled for other time periods.

2.2.2. Induction Period


The mass conservation equation for this period is expressed as:

The rate of reaction R is assumed to have the same driving force that Natarajen, et al.
(1994) proposed:

where fe is the fugacity of the gas at the three-phase equilibrium condition, n induction is the
nucleation index, and kinduction is a constant. The reaction index, ninduction is not assumed to
be equal to one. Rather, its value is determined by means of the experimental data. The

L f gV
departure from equilibrium can be expressed by either f g or These ratios will be
fe
f gV
equal to unity when equilibrium is attained. The ratio can be estimated using an
fe

equation of state for the gas phase. The water vapor pressure can be neglected and the

f gL
gas phase is assumed pure methane. Further, can be estimated using an equation of
fe

state for the liquid system, but the mole fraction of the gas in the liquid cannot be
determined at non-equilibrium conditions. The concentration of the gas in the liquid will
vary with the distance from the interface and this variation will be related to the mixing
rate, among other factors. Therefore, there are many factors that can interfere with the

f gL
calculation adding errors to the expression of the driving force and the reaction
fe

rate. Therefore, the driving force used in the non-equilibrium model calculations is

f gV
taken as (  1)
fe

2.2.3. Growth Period


The start of this period was identified by an abrupt change in the slope of the cumulative
gas consumption curve. This coincides with the turbidity point observations of
Natarajan et al. [8]. Thus, the growth period definition for this model is similar to that
proposed by Natarajan et al. Similar to the induction period, it is assumed that the gas
consumed from the gas phase is totally converted into hydrate during the growth period.
The gas consumed from the liquid by the reaction will be replaced, and the saturation of
the liquid will hold during this period. Because we assume that all hydrate being formed
migrates rapidly to the gas-liquid interface and grow as a “layer” there, we can compute
the “thickness” of this “layer” using

where R is the rate of reaction, and is the molar gas density. The kinetics of
this reaction will be assumed to follow a power-law form as:
The fugacity of the gas is a function of the temperature and pressure at the reaction
condition. The reaction is assumed to occur below the hydrate layer as soon as the gas
permeating through the hydrate layer reaches the water below. The pressure below the
hydrate layer is assumed to be equal to the fugacity of the gas and is calculated using
Darcy’s law for gas flow through porous media (Collins, [13]):

where b is the Klinkenberg constant. Because pressures in the experiments are typically
fairly high, the term involving the Klinkenberg constant is neglected. As for the
remaining terms, P1 is the pressure at the gas-hydrate interface, P2 is the pressure at the
liquid-hydrate interface, KH is the effective permeability of the hydrate patches, and  g
is the gas viscosity. The pressure at the liquid-hydrate interface, P2 is assumed to be
equal to the fugacity of the gas. Note that the assumption of a Darcy flux is only for
convenience. Any closed form equation that would allow the calculation of the pressure
below the hydrate “layer” could be used.

3. EXPERIMENTAL STUDIES
A number of publications present data on hydrate formation and/or dissociation.
Experimental equipment and procedures differ slightly but are essentially similar. As
this paper analyzes the in-house data from experiments at NETL, these experiments will
be described in detail.
Experiments were conducted in a high-pressure view cell (Figure 4). The cell is
constructed of 316-stainless steel with 6.35 cm (2.5 inches) OD and is 27.4 cm
(11 inches) in length. The internal volume of the cell is approximately 40 ml. The cell is
fitted with two machined end-caps, one of which contains a sapphire window to allow
for observation of the contents of the cell using a CCD camera. The cell is fitted with
ports to accommodate the fill gas inlet and reaction product outlet, a pressure transducer
to monitor the internal pressure of the gas inside the cell, and a thermocouple that
terminates inside the cavity of the cell to monitor the temperature of the liquid/hydrate
mixture.
While the working pressure of the cell is rated at 220 MPa (32,000 psia), all
experiments were conducted at 13.8 MPa (2,000 psig) or less. The temperature of the
cell is controlled by the flow of a glycol/water solution from an external circulating
temperature bath through a coil of 0.64 cm (¼ inch) copper tubing that is wrapped
around the outside of the cell. Several layers of insulating material are wrapped around
the cell to help maintain constant temperature.
A typical experiment involves filling the cell with approximately 20 ml (±0.1 ml) of
double-distilled water. A Teflon®-coated stir bar is also placed inside the cell. The end
caps are placed on the cell and tightened to specifications. An external magnetic stirrer
is used to obtain a high degree of vortex mixing inside the cell. The cell is connected to
a gas manifold and purged several times with methane. Following the purge procedure,
the cell is charged with methane at pressures of 8.3–10.3 MPa (1,200–1,500 psig).
Using the external circulating temperature bath, the temperature of the water in the cell
is gradually lowered until formation of the methane hydrate is observed. Then, the
temperature of the cell is lowered to approximately –10° C and held constant as the
unabsorbed methane is released from

the cell. The cell is finally allowed to warm slowly to a temperature between 10° C and
room temperature, which causes the hydrate to release its absorbed methane gas.

3.1. Application of the Equilibrium Approach


An example plot of the formation and dissociation results is shown in Figure 5 and the
slope values for all of the runs are shown in Table 1. The slope differences between the
formation and dissociation data are pronounced, while the slopes of the dissociation
data and the predicted values are similar. The predicted values were calculated
assuming that the hydrate formation occurs at equilibrium condition. Thus, we can
deduce that the dissociation occurs at equilibrium conditions while the formation does
not.

Figure 5. Experimental and Software Formation and Dissociation Trend Lines for Run 2.
4. CONCLUSIONS
The proposed hydrate formation and dissociation models describe fairly well the
behavior of the experimental data for different experimental conditions and apparatus,
while providing important insights into the mechanism of the hydrate formation and
dissociation processes.

You might also like