Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Fluid Phase Equilibria 388 (2015) 160–168

Contents lists available at ScienceDirect

Fluid Phase Equilibria


j o u r n a l h o m e p a g e : w w w . e l s e v i e r . c o m / l o c a t e / fl u i d

Phase equilibria, solubility and modeling study of CO2/CH4 + tetra-n-


butylammonium bromide aqueous semi-clathrate systems
Jonathan Verrett, Jean-Sébastien Renault-Crispo, Phillip Servio *
Department of Chemical Engineering, McGill University, Canada

A R T I C L E I N F O A B S T R A C T

Article history: Interest has grown in tetra-n-butylammonium bromide (TBAB) semi-clathrates due to their formation at
Received 31 March 2014 lower pressures and higher temperatures than conventional gas hydrates. This study focuses on TBAB
Received in revised form 19 December 2014 semi-clathrates formed with carbon dioxide and methane and is to our knowledge the first study to
Accepted 29 December 2014
report and model semi-clathrate former solubility under hydrate–liquid–vapor equilibrium conditions. A
Available online 2 January 2015
thermodynamic model was developed using the Trebble–Bishnoi equation of state to calculate vapor and
liquid phase fugacity. Liquid activity was determined using the electrolyte non-random two liquid model
Keywords:
(eNRTL). The solid semi-clathrate phase was modeled using a modified van der Waals–Platteeuw model
Semi-clathrate hydrate
Phase equilibria
(vdW–P), with parameters re-optimized based on the data obtained in this study. Equilibrium conditions
Solubility were successfully modeled over the temperature range of 281–294 K, pressure range of 377–11,000 kPa
TBAB and TBAB composition range of 5–40 wt%. The model functioned well for carbon dioxide with an average
Modeling absolute relative error (AARE) of 4.7% for pressure prediction and an AARE of 17.5% for solubility. Methane
was more difficult to model and yielded AAREs of 21.6% for pressure and 32.5% for solubility. Further
understanding of semi-clathrate crystal structure and interactions between methane and TBAB would
aid in improving the model.
ã 2014 Elsevier B.V. All rights reserved.

1. Introduction means. One method uses chemical promoters added to the


solution to increase the rate of hydrate formation. These can
Clathrate hydrates are non-stoichiometric crystalline be separated into two broad categories of either kinetic or
compounds composed of a water lattice held together through thermodynamic promoters. Kinetic promotion involves substances
hydrogen bonding that are further stabilized by guest molecules that have been shown to significantly increase hydrate formation
occupying cavities in the lattice structure [1]. Clathrates typically rates without affecting thermodynamic equilibrium. Commonly
form at moderate temperatures (<300 K) and high pressures studied kinetic promoters are generally surfactants, the most
(>1 MPa), although formation conditions vary greatly based on the notable of which is sodium dodecyl sulfate [4,5]. Thermodynamic
guest species. Industrial interest in gas hydrates was initially promotion involves the addition of another guest substance that
spurred on in the 1930’s following a report by Hammerschmidt makes hydrate formation more energetically favorable. This shifts
citing hydrate formation as the reason for blockages in natural gas equilibrium conditions and either reduces the pressure or
pipelines [2]. Since this time, many applications have been correspondingly increases the temperature at which hydrates will
proposed for these compounds such as gas transport, storage form. Examples of these thermodynamic promoters include
and separation as well as water desalination and food processing tetrahydrofuran (THF) and cyclopentane, which generally integrate
[3]. Though energetically favorable for use in many applications, into larger cages in the hydrate structure and thus facilitate the
slow formation kinetics as well as lack of economic and scalability storage of smaller gas molecules such as methane (CH4) or carbon
studies have prevented larger scale use of hydrate technologies [3]. dioxide (CO2) in the empty smaller cages [6–8].
To address the slow kinetics of formation of gas hydrates, Promoter research has recently shifted to hydrate compounds
research has been undertaken to promote growth using a variety of known as semi-clathrates. The lattice structure of these compounds
is different from those of conventional clathrates that contain only
water and guest molecules. Semi-clathrates integrate organic salts
* Corresponding author at: Room 3060, Wong Building, 3610 University Street,
into their crystal structure, thereby further stabilizing the crystal
Montréal, Québec H3A 0C5, Canada. Tel.: +1 5143981026; fax: +1 5143986678. lattice and allowing formation at more moderate conditions
E-mail address: phillip.servio@mcgill.ca (P. Servio). (lower pressures and/or higher temperatures) [9,10]. Studies have

http://dx.doi.org/10.1016/j.fluid.2014.12.045
0378-3812/ ã 2014 Elsevier B.V. All rights reserved.
J. Verrett et al. / Fluid Phase Equilibria 388 (2015) 160–168 161

modeled the solubility, being the liquid mole fraction at


Nomenclature
equilibrium, of guest gases in the liquid under three phase
hydrate–liquid–gas equilibrium. Knowledge of liquid phase
AARE Average absolute relative error
composition, most notably with respect to the guest gas, is
C Langmuir constant (Pa1)
essential for developing kinetic models for reactor design [18]. The
f Fugacity (Pa)
importance of accurate modeling of the liquid phase has been
g Objective function
highlighted in literature showing that equilibrium values obtained
g Activity coefficient
by thermodynamic models are much more sensitive to liquid
m Chemical potential (J/mol)
parameters than to gas parameters [19]. To enhance understanding
N Hydration number
of the semi-clathrate equilibrium, the current study investigates
NA Avogadro’s number
equilibrium pressure, temperature and liquid phase composition
n Number of cavities per water molecule
of TBAB hydrates formed with carbon dioxide or methane. A
P Pressure (Pa)
thermodynamic model is developed to accurately predict three
R Universal gas constant (J/mol K)
phase semi-clathrate equilibrium conditions. The model consists
r Density (mol/m3)
of the Trebble–Bishnoi equation of state for calculating the gas
s Standard deviation
fugacity, the electrolyte NRTL (eNRTL) model to calculate liquid
T Temperature (K)
activity and the modified vdW–P model for hydrate phase energy.
u Uncertainty
This is the first study to measure and model liquid solubility for a
V Molar volume (m3/mol)
three phase hydrate–liquid–gas semi-clathrate system.
x Mole fraction
y General variable
2. Experimental
wp TBAB weight fraction
2.1. Experimental setup
Subscripts
j Component j
A detailed description of the experimental setup can be found in a
i Cage type i
previous report [20]. Briefly it consists of a 600-cm3 stainless steel
TBAB Tetra-n-butylammonium bromide
reactor with a mounted stirrer and viewing windows. The setup
w Water
provides access to liquid and gas samples via three ports at the top,
middle and bottom of the reactor. Reactor temperature is controlled
Superscripts
through a glycol water bath. Pressure measurements are performed
H Hydrate
using a Rosemount transducer with a span of 0–14 MPa and an
L Liquid
accuracy of 0.065% of the span. Temperatures are monitored using
V Vapor
Omega platinum resistance temperature device probes.
b Empty clathrate
p Phase pi 2.2. Materials
sat Saturation
Tetra-n-butylammonium bromide (TBAB) was purchased from
Sigma–Aldrich as a 50 weight percent (wt%) aqueous solution. All
focused on using variety of halide salts as promoters to form gases used were obtained from MEGS Inc., and included ultra-high
semi-clathrates. One of the most extensively studied salts to date is purity methane gas (99.99%) and carbon dioxide gas (99.99%).
tetra-n-butylammonium bromide (TBAB), which has been very
successful at forming hydrates at more moderate conditions [11]. 2.3. Procedure for equilibrium experiments
Application of semi-clathrates requires knowledge of their
thermodynamic properties and phase equilibria. Previous studies The reactor contains a holdup volume of roughly 40 mL and was
have focused on measuring pressure, temperature and TBAB liquid cleaned by washing three times with 360 mL of the desired TBAB
concentration at equilibrium for semi-clathrates with single and solution. Following this, 360 mL of the test solution was loaded into
multiple guest gases [12,13]. Thermodynamic equations of state the reactor. The reactor gas was purged and replaced with the
and correlations have been modified to predict equilibrium desired gas by pressurizing to 1000 kPa, allowing time to mix, and
temperature and pressure for semi-clathrates while taking into then depressurizing to 100 kPa. This procedure to replace the gas
account the TBAB composition in the system [14]. These models was repeated three times following which the pressure was
have used a variety of equations to describe gas phase fugacities, brought up to a level greater than the estimated equilibrium
notably the Patel–Teja (PT) [12] and Soave–Redlich–Kwong (SRK) pressure from previous literature data [9,10,21–24]. Stirring was
[15] equations of state. Liquid phase modeling has been performed started and once hydrates formed, the pressure was decreased to
using techniques such as the Non-Random Two-Liquid (NRTL) [16], be near the estimated equilibrium pressure. The system was then
Statistical Associating Fluid Theory with Range of Variable left to equilibrate under stirring for a minimum of 12 h. During this
Electrolytes (SAFT–VRE) [17], as well as correlations based on time, pressure was monitored and changes of no more than 1 kPa
osmotic activity [12]. The hydrate phase has been modeled in the were observed within the hour before measurement.
literature using two methods to account for the effect of TBAB in At the time of measurement, the pressure and temperature
the semi-clathrate structure. The first is a modified version of the were recorded and the stirring was turned off. Hydrates would
van der Waals–Platteeuw model (vdW–P) [16] and the second is a agglomerate and float or sink based on their guest compound. Five
modified two-step formation model proposed by Chen and Guo liquid samples of roughly 10 mL were then drawn into sample
[15]. This shows that a variety of thermodynamic models have bombs through a high-pressure inline filter with a 20-nm nominal
been adapted to accurately predict equilibrium pressure and rating (Norman Filters). The sample bombs used had been
temperature of three phase semi-clathrate systems. weighed, vacuum pumped and put to chill in the reactor bath
There has been substantial research progress in the past decade before sampling to ensure thermal consistency. The first liquid
on semi-clathrates, however studies have not yet measured and sample taken was used to clear the sample line and was then
162 J. Verrett et al. / Fluid Phase Equilibria 388 (2015) 160–168
[(Fig._1)TD$IG]

Fig. 1. A schematic of the procedure for determining the compositions of the liquid sample taken from the crystallizer.
!
bH
discarded, while the other four samples were used for analysis. A H b Dmw ðT; PÞ
f w ðT; PÞ ¼ f w ðT; PÞexp (2)
schematic showing the analysis of each sample bomb can be seen RT
in Fig. 1. This consisted of first weighing of the bombs following b
sample loading. A gasometer was then used to depressurize the where f w is the fugacity in the hypothetical empty clathrate phase
bH
sample bombs, which were left for 2 h to equilibrate. Gas volume, and Dmw is the chemical potential difference between the filled
temperature, and atmospheric pressure were noted. The liquid was and empty hydrate phases as defined below by van der Waals and
then taken out of the sample bombs, weighed, and placed on a Platteeuw [28]:
hotplate to evaporate the water and leave only TBAB salt. Following 0 1ni
2 h at 80  C, the water had completely evaporated leaving only the DmwbH ðT; PÞ X NS X
NH
 ¼ ln@1 þ C i;j ðT Þf j ðT; PÞA (3)
solid TBAB to be weighed. The samples bombs were then washed RT i¼1 j¼1
and the process repeated at the next measurement condition.
where f j is the fugacity of hydrate former j,vi is the number of
3. Vapor–liquid–hydrate equilibrium model cavities of type i per water molecule and Ci,j is a Langmuir constant
given by a correlation proposed by Parrish and Prausnitz [27] that
The basic thermodynamic equation for the three-phase vapor– is valid in the range 260–300 K.
liquid–hydrate equilibrium is fugacity equality in all phases for The fugacity of water in the hypothetical empty hydrate lattice
each component, as shown in Eq. (1): and in the liquid is given by the following two common
H L V
thermodynamic equations:
f i ðT; PÞ ¼ f i ðT; PÞ ¼ f i ðT; PÞ½i ¼ 1; . . . ; N (1) !
p b b V bw ðT; PÞ½P  Psat;b
ðTÞ
where N is the total number of components in the system and f i is f w ðT; PÞ ¼ Psat;
w ðTÞexp w
(4)
RT
the fugacity of component i in phase p.
p
The fugacity of water and gas components in the liquid and where Psat;w is the saturation pressure of water and V p w is the molar
vapor phase can be calculated from a suitable equation of state. In volume of water in phase p, which in the case of Eq. (4) is the
this study, the Trebble–Bishnoi equation of state is used since it has empty hydrate lattice (b), and
been shown to accurately model interactions between the water !
and gasses chosen in this study [25,26]. Under the temperature and V Lw ðT; PÞ½P  Psat;L
w ðTÞ
f w ðT; PÞ ¼ xw ðT; PÞg w ðxw ; TÞPw ðTÞexp
L sat;L
pressure conditions of the study other equations of state such as RT
the Patel–Teja would likely give similar results. It was assumed that
(5)
no TBAB is present in the gas phase because of its non-volatile
nature. The critical properties and interaction parameters used in where xw is the mole fraction of water in the liquid phase and g w is
the equation of state are shown in Tables 1 and 2, respectively. the activity coefficient of water. The fugacity of the TBAB salt is
Additionally, the solubility of the gas in the liquid phase is assumed determined using a similar basic thermodynamic formula [16]
to be unaffected by the presence of TBAB, which is further shown as Eq. (6).
discussed in Section 4.3 of this paper. L
f TBAB ðT; PÞ
To calculate water fugacity in the hydrate phase, Parrish and
0 L h i1
Prausnitz [27] proposed a modification to the vdW–P [28] model to V TBAB ðT; PÞ P  Psat;L
TBAB ðT Þ
¼ g
xLTBAB ðT; PÞ TBAB ðxw ; TÞPsat;L @ A
TBAB ðT Þexp
calculate the chemical potential of water in the hydrate phase.
Klauda and Sandler [29] calculated water fugacity using the RT
chemical potential from Parrish and Prausnitz [27] and this same (6)
technique is used in this paper. Eqs. (2)–(5) were taken from
Klauda and Sandler [29]:

Table 1
Critical parameters of different compounds for the Trebble–Bishnoi equation of state [25,26].

Component Tc (K) Pc (MPa) v dc (m3/mol) zc q1 q2


H2O 647.286 22.09 0.34375 18.03 0.24403 0.46195 0.23002
CO2 304.21 7.3825 0.225 33.82 0.28744 0.44277 0.2250
CH4 190.555 4.595 0.01001 25.64 0.31135 0.36438 0.05927
J. Verrett et al. / Fluid Phase Equilibria 388 (2015) 160–168 163

Table 2 Table 3
Interaction parameters between water–gas for the Trebble–Bishnoi equation of Constants in Eqs. (14) and (15) for calculating TBAB density [38].
state [27].
Subscript si ri qi
Components KA KB KC KD
1 1.707  108 5.693  106 4.549  104
CO2–H2O 0.9688 0.5181 0.3757 0.1647 2 4.570  109 3.099  106 5.304  104
CH4–H2O 0.4199 0.1727 0.0001 1.2274 3 0 4.088  108 7.091 106

 
The activity coefficients of both water and TBAB were 6 1353:108782
Psat
TBAB ½Pa ¼ 10 exp 0:999694  (11)
determined using the electrolyte NRTL (eNRTL) activity T½K
coefficient model [30,31]. In this model, the non-randomness
factor was set to a = 0.2 [30] and the binary parameters used were The molar volume equations are from Klauda and Sandler [29].
t H2 O;TBAB ¼ 8:3169 and t TBAB;H2 O ¼ 3:6717 [14,32]. The empty clathrate molar volume constants were obtained by
The objective function Eq. (7) can be found by putting together regressing the experimental thermal expansion of hydrate data by
Eqs. (1)–(6) Tse [33].
gðT; PÞ ¼ 0  3
m
V bw
 b mol
sat;b 
Psat; b
exp V w ½PP w    ni 
¼1
w
 L
RT
 P 1 þ S C i;j f j 1030 NA
sat;L
V w ½PPw 
(7) ¼ ð11:835 þ 2:217  105 T½K þ 2:242  106 T½K2 Þ3
xw g w Psat;L
w exp RT
i j 46
8:006  109 P½MPa þ 5:448  1012 P½MPa2 (12)
where i, found in the product term, denotes the different types of where NA is Avogadro’s number and P is in units of MPa.
cavities in the hydrate structure. As will be explained in more detail The molar volume of pure water equation was obtained using
later, the TBAB–water system forms a structure with two large experimental data from Perry’s Handbook [34] and NIST [35].
cages occupied by TBAB and one small cage occupied by a guest gas  3
m 
molecule. The summation of all j’s is over each guest molecule. For V Lw ¼ exp  10:9241 þ 2:5  104ðT½K273:15Þ  3532
every specific cavity type, there is only one possible guest and as a mol
result, the objective function becomes:
104 ðP½MPa  0:101325Þ þ 1:559
gðT; PÞ ¼ 0 
  107 ðP½MPa  0:101325Þ2 (13)
b sat;b
b V w ½PPw 
Psat;
w exp
½ð1 þ C large1;TBAB f TBAB Þnlarge1
RT
¼1 
V Lw ½PPsat;L
w 
xw g sat;L
w P w exp RT The density, r, of the TBAB aqueous solution (inverse of molar
nlarge2 nsmall volume, V LTBAB ) was obtained from the Eqs. (14) and (15), which are
ð1 þ C large2;TBAB f TBAB Þ  ð1 þ C small;gas f gas Þ  (8)
both a function of temperature and TBAB weight fraction in water
[36,37]. The constants qi, ri and si are shown in Table 3.
The method for solving the three phase equilibrium using  
kg   
Eq. (8) is as follows. At a fixed temperature, the objective function rTBAB 3 ¼ rw þ o1 100wp þ o2 100wp 2 þ o3 100wp 3 (14)
m
is solved for a pressure using the bisection method. One must know
the pressure range to initialize the problem. Following this, the
temperature is changed and the process is repeated for the whole
oi ¼ q i þ r i T þ s i T 2 (15)
temperature range. The different model parameters and equations
used are explained below.
The first parameter is the saturation pressure of water in the The method of obtaining the Langmuir constants is derived
empty clathrate lattice. Eslamimanesh et al. [16] proposed the from Parrish and Prausnitz [27]. Eslamimanesh et al. [16] modified
following equation that takes into account the effect of TBAB the Langmuir constant equation to take into account the TBAB
weight percent, wP, on the cage structure by adding the adjustable presence in the hydrate lattice. The TBA+ molecules take up two
parameter a large cages in the hydrate lattice (larges 1 and 2), while each gas
  molecule takes up a small cage similar to the small cage of
6003:9
Pbw ½Pa ¼ 10000exp 17:44  þ a  wp (9) structure I clathrate. The affinity of the gas to stay in the hydrate
T½K
lattice is determined by this Langmuir constant and is assumed to
be the same as for structure I small cages without TBAB present
The saturation pressure of pure water is given by the following (similar water–gas interactions are assumed). For this reason, the
equation [29]: Langmuir constants for the gas in the small cages were taken
 directly from Parrish and Prausnitz [27] and can be seen in Table 4.
5500:9332
Psat
w ½Pa ¼ exp 4:1539lnðT½KÞ  The Langmuir constants for TBAB are found by adding a TBAB
T½K
weight fraction variable into the original equations as shown in
þ7:6537  16:1277  103 T½KÞ (10)
Table 4
Langmuir constants in Eq. (18) for methane and carbon dioxide [28].
An equation to calculate the saturation pressure of TBAB
Component D E
was taken from a paper by Eslamimanesh et al. [16]. The authors
CO2 0.0011978 2860.5
did not mention the constants in the publication but by author
CH4 0.0037237 2708.8
correspondence Eq. (11) was obtained.
164 J. Verrett et al. / Fluid Phase Equilibria 388 (2015) 160–168
[(Fig._2)TD$IG]
Eqs. (16) and (17).
  c
b
C large1 ¼ exp ð1 þ d  wp Þ (16)
T T

e g 
C large2 ¼ exp ð1 þ h  wp Þ (17)
T T

   
D E
C small ¼ exp (18)
T T

The number of cages per water molecule in a unit hydrate cell


can be determined from the following three equations:
4
nlarge1
typeA ¼ b
(19)
NW

4
nlarge2
typeA ¼ b
(20) Fig. 2. Experimental (vapor + liquid + hydrate) equilibrium pressures at various
NW temperatures and TBAB loading concentrations for the system CO2–TBAB–H2O. This
work: ", 5 wt%; *, 10 wt%; v, 40 wt%; literature: [TD$INLE], 5 wt% [21]; [TD$INLE], 5 wt% [23]; [TD$INLE],
10 wt% [21]; [TD$INLE], 10 wt% [23]; [TD$INLE], 40 wt% [24]; [TD$INLE], 42.7 wt% [9]; model: [TD$INLE] 5 wt%; [TD$INLE]
10 wt%; [TD$INLE] 40 wt%.
6
nsmall
typeA ¼ b
(21)
NW

different equations of state and activity coefficient models) as


b
The number of water molecules per hydrate cell, NW , is double well as different equations for molar volumes and saturation
the value of what is called the hydration number. The hydration pressures. Therefore, these constants required re-optimization
number for TBAB–water hydrates varies greatly in the literature using the hydrate experimental data. The parameter optimization
with values between 24 and 38 being reported [38–41]. Pressure, was performed using the pressure–temperature equilibrium
temperature, TBAB weight fraction and guests gases present all experimental data reported in this publication. A stochastic
have an effect on the final hydrate lattice structure. Some reports method was used to find the minima because of the large number
state the presence of two different structures, named by Shimada of constants and high non-linearity of the model. The optimized
et al. [41] type A (hydration number of 26) and type B (hydration parameters are reported in Table 5. Note that the methane
number of 38). Tests on TBAB hydrate structure under methane at hydration number ends up being optimized to the hydration
atmospheric pressure have shown that type B is more stable below number of type B hydrate, consistent with the results of previous
approximately 18 wt% and type A is more stable above this studies showing this structure to be more stable than type A
concentration [42]; similar experiments in air show a similar under roughly 20 wt% [38,42].
transition point of roughly 20 wt% [38]. In this paper, the number of
water molecules per hydrate cell has been taken as a variable in the [(Fig._3)TD$IG]
model. This value has been optimized for each guest gas and
therefore has been fixed for any value of temperature, pressure and
TBAB weight fraction with the same guest gas. This assumption
was made to reduce the amount of fitting parameters in the model.
The validity and possible effect of this assumption is discussed in
Section 4.3.
Many constants are required to be determined to complete the
model. These constants were already found by optimization by
Eslamimanesh et al. [16] but did not give adequate results with
the proposed model. Many factors can account for the ineffective
constants, such as differences in the baseline for fugacity (from

Table 5
Optimized parameters for the proposed model.

Component CO2 CH4


b 28 38
Hydration number NW
a 0.2575 0.0189
b 0.6462 0.5185
c 3408.6 2629.2
d 3.6414 -0.8745 Fig. 3. Experimental (vapor + liquid + hydrate) equilibrium pressures at various
e 0.1164 0.0065 temperatures and TBAB loading concentrations for the system CH4–TBAB–H2O. This
g 7202.4 8534.0 work: ", 5 wt%; *, 10 wt%; v, 20 wt%; literature: [TD$INLE], 5 wt% [10]; [TD$INLE], 5 wt% [21]; [TD$INLE]
h 0.6669 0.1145 10 wt% [10]; [TD$INLE], 9.9 wt% [22]; [TD$INLE], 20 wt% [9]; [TD$INLE], 19.7 wt% [22]; model: [TD$INLE] 5 wt%; [TD$INLE]
10 wt%; [TD$INLE] 20 wt%.
J. Verrett et al. / Fluid Phase Equilibria 388 (2015) 160–168 165

Table 6
Experimental (vapor + liquid + hydrate) equilibrium data for temperature T, pressure P, carbon dioxide mole fraction xCO2 and TBAB mole fraction xTBAB with standard
uncertainties u(xCO2) and u(xTBAB) for the system CO2–TBAB–H2O.a

TBAB loading concentration/wt% T/K P/MPa xCO2 [x103] u(xCO2) [x103] xTBAB [x103] u(xTBAB) [x103]
5 281.1 0.859 7.054 0.038 2.436 0.017
5 283.1 1.329 10.007 0.091 2.714 0.028
5 285.1 2.526 16.470 0.046 2.635 0.016
5 286.1 3.254 19.474 0.026 2.729 0.004
5 287.2 4.678 23.890 0.213 2.711 0.033
10 283.1 0.683 4.970 0.236 5.062 0.186
10 285.1 1.352 9.080 0.090 4.939 0.011
10 287.4 2.317 13.330 0.041 5.475 0.032
10 288.2 3.011 16.193 0.074 5.526 0.100
10 289.1 3.975 19.183 0.030 5.471 0.035
40 285.1 0.477 3.144 0.050 35.577 0.368
40 287.1 1.021 9.462 0.564 36.180 0.372
40 289.2 1.957 13.411 0.374 37.537 0.432
40 290.1 2.609 15.535 0.115 34.310 0.392
40 291.2 3.594 21.418 0.753 36.157 0.317
a
Standard uncertainties are u(T) = 0.1 K and u(P) = 0.009 MPa.

computed taking into account the mass of guest and water in the
4. Results and discussion vapor using Eq. (23).

4.1. Solubility calculations mLo ¼ mo  mVg  mVw (23)

The solubility calculations performed are similar to those done


on previous hydrate systems with a slight modification to The fraction of TBAB in the liquid phase was computed using the
determine the TBAB content in the liquid [43]. As shown by measured weights of solution before and after drying. This was
the measurement procedure in Fig. 1, it was assumed the gas then used to find the TBAB and water mass fractions in the liquid
recovered in the gasometer contained only water vapor and guest shown in Eqs. (24) and (25).
gas (either carbon dioxide or methane). Gas species were mLt ¼ mLo  xt (24)
assumed to be ideal at ambient conditions and the mass of each
species, i, was calculated using Eq. (22) below. Water pressure was and
assumed to be the vapor pressure at the ambient temperature and mLw ¼ mLo  ð1  xt Þ (25)
was calculated using the Antoine parameters found in the NIST
webbook [44]. Guest gas pressure was then calculated using the
ambient pressure as total pressure and subtracting the water vapor The fraction of guest gas remaining in the liquid was computed
pressure. using the appropriate Henry’s constants found in the NIST
webbook [44]. This value represented less than 0.1% of the total
Pi V gasometer MW i
mVi ¼ (22) number of moles of liquid water, and was thus not taken into
RT ambient account in calculating the number of moles of liquid water.
However it can represent up to 4% of the total number of moles
The weight of the liquid following depressurization was
[(Fig._4)TD$IG] [(Fig._5)TD$IG]

Fig. 4. Experimental (vapor + liquid + hydrate) equilibrium CO2 solubilities Fig. 5. Experimental (vapor + liquid + hydrate) equilibrium CH4 solubilities
at various temperatures and TBAB loading concentrations for the system at various temperatures and TBAB loading concentrations for the system
CO2–TBAB–H2O. This work: ", 5 wt%; [TD$INLE], 10 wt%; [TD$INLE], 40 wt%; model: [TD$INLE] 5 wt%; [TD$INLE] CH4–TBAB–H2O. This work: ", 5 wt%; [TD$INLE], 10 wt%; [TD$INLE], 20 wt%; model: [TD$INLE] 5 wt%; [TD$INLE]
10 wt%; [TD$INLE] 40 wt%. 10 wt%; [TD$INLE] 20 wt%.
166 J. Verrett et al. / Fluid Phase Equilibria 388 (2015) 160–168

of the guest gas and is thus important to include in calculating values of temperature, pressure and liquid mole fractions can be
the total number of moles of guest gas in the sample. found in Table 6 for carbon dioxide and Table 7 for methane.
Standard uncertainties for temperature and pressure were
estimated to be u(T) = 0.1 K and u(P) = 0.009 MPa respectively. 4.3. Model fit
Uncertainty values for mole fractions were obtained using the
following formulas [45]. The solubility of the guest gas molecule in the liquid phase was
s modeled successfully for the TBAB–H2O–CO2 and TBAB–H2O–CH4
u ¼ pffiffiffi (26) equilibrium systems with gas, liquid and hydrate phases present.
n
The model temperature ranged from 281 K to 294 K and the
pressure ranged from 377 kPa to 11,000 kPa. Fig. 2 shows the
and pressure–temperature three-phase equilibrium line for the
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Pn ffi TBAB–water with carbon dioxide as a guest gas. The Average
k¼1 ðyk  yÞ Absolute Relative Error (AARE) is used as a measure of model fit
s¼ (27)
ðn  1Þ with the experimental data and is shown for a general variable y in
Eq. (28):
where y is a general variable, with y being the average of a given set
100 X yexp  ymodel
of n values of y, s is the standard deviation and u is the standard ndata
k
uncertainty. % AARE ¼ j k j (28)
ndata k¼1 yexp
k

4.2. Equilibrium measurements


For CO2, the AARE of pressure compared to the experimental
Equilibrium pressure and temperature data for various TBAB data collected in this publication is 4.7%. The model fits the CO2
loading concentrations can be found in Fig. 2 for carbon dioxide data well and follows a trend similar to the experimental data.
and Fig. 3 for methane. Previous literature values for temperature With methane, the parameter fitting was not as successful. The
and pressure at equilibrium at similar TBAB concentrations are model, seen in Fig. 3, has an AARE of pressure of 21.6%. Liquid mole
shown for comparison. The pressure values found in this study are fraction predictions show similar fit for the two gases, with the
similar or generally higher than those found elsewhere for the methane model being less accurate than the CO2 model. The model
same temperature. Higher pressures values are most likely for liquid mole fraction of CO2 shown in Fig. 4 has an AARE of 17.5%.
found because previous literature studies measure the point of For methane in Fig. 5 the model yields an AARE of solubility of
hydrate dissociation using the isochoric search method to find the 32.5%.
three-phase equilibrium. With this method, the amount of TBAB A variety of reasons may account for the lack of fit of the
in the liquid phase is roughly equal to that of the loading proposed model. Assumptions made in the model could have a
composition. In this study, however, the hydrate phase was significant impact on the output results. The assumption that
included during measurement. The TBAB loading concentrations TBAB does not affect the two phase flash (and therefore the
used were generally lower than the stoichiometric amount of solubility value) is thought to be a source of error. This
TBAB in the semi-clathrate structure, which is estimated to be assumption has a more pronounced effect on the methane model
near 40 wt%, however this depends on the crystal structure [42]. because its concentration is considerably lower than carbon
The amount of TBAB in the liquid is therefore expected to be lower dioxide. Thus a small difference in the methane liquid mole
than the loading composition because a larger concentration of fraction (and fugacity) affects the three-phase equilibrium much
TBAB is stored in the solid hydrate phase. Indeed, liquid TBAB more than the same difference in the carbon dioxide mole
measurements, given in Table 6, showed compositions slightly fraction. For example, if the temperature and pressure values
lower than the loading compositions. With less TBAB in the were fixed to the experimental values obtained and used to
solution, the promotion effect is less pronounced and it is calculate the vapor–liquid equilibrium, the values computed for
expected that the equilibrium data would shift to higher the methane liquid mole fraction would yield an AARE of 16.0%.
pressures or lower temperatures. This is considerably lower than the AARE of 32.5% obtained with
Liquid solubility measurements for carbon dioxide and the three-phase model and shows the sensitivity of such models
methane systems can be found in Figs. 4 and 5, respectively. Exact to liquid phase compositions. The validity of this assumption can

Table 7
Experimental (vapor + liquid + hydrate) equilibrium data for temperature T, pressure P, methane mole fraction xCH4 and TBAB mole fraction xTBAB with standard uncertainties
u(xCH4) and u(xTBAB) for the system CH4–TBAB–H2O.a

TBAB loading concentration/wt% T/K P/MPa xCH4 [x103] u(xCH4) [x103] xTBAB [x103] u(xTBAB) [x103]
5 283.3 2.417 0.700 0.011 2.208 0.036
5 285.1 3.496 0.880 0.022 2.847 0.044
5 287.1 5.137 1.180 0.003 2.770 0.012
5 289.2 8.580 1.761 0.014 2.884 0.052
5 290.2 11.093 2.037 0.003 2.783 0.024
10 283.2 1.047 0.293 0.007 4.666 0.027
10 285.2 1.652 0.380 0.009 5.365 0.092
10 287.1 2.904 0.700 0.003 5.333 0.032
10 289.3 4.733 0.868 0.006 5.623 0.021
10 291.2 8.136 1.523 0.001 5.380 0.049
20 287.2 1.435 0.296 0.011 11.028 0.037
20 289.1 3.078 0.648 0.004 10.898 0.042
20 291.1 5.434 0.929 0.002 11.075 0.045
20 293.2 8.202 1.378 0.011 11.025 0.051
20 294.2 11.013 1.693 0.005 10.847 0.038
a
Standard uncertainties are u(T) = 0.1 K and u(P) = 0.009 MPa.
J. Verrett et al. / Fluid Phase Equilibria 388 (2015) 160–168 167

Table 8
AAREs for 10% change in the empty cage molar volume parameter.

% Change of empty cage molar volume AARE of CO2/% AARE of CH4/%


0 Pressure 4.7 21.6
0 Liquid mole fraction 17.5 32.5
+ 10 Pressure 5.7 23.0
+ 10 Liquid mole fraction 19.1 35.8
 10 Pressure 5.8 21.7
 10 Liquid mole fraction 16.1 29.9

also be assessed using solubility data from previous studies in of 377–11,000 kPa and TBAB composition range of 5–40 wt%. When
comparison to the data found in this paper. The closest data point compared to experimental data, the equilibrium pressure was
for methane solubility found from literature was at 283.37 K and predicted with an AARE of 4.7% for CO2 and 21.6% for CH4.
2.799 MPa, where the solubility in pure water was measured to be Equilibrium gas solubility values were predicted with an AARE of
0.000893 [46]. The data point reported in this paper in Table 7 is 17.5% for CO2 and 32.5% for CH4. Significantly larger error for the
at 283.3 K and 2.321 MPa with 5 wt% TBAB loading where the methane model is attributed to its more complex interactions with
methane’s solubility is 0.000700. Another example using CO2 TBAB–water and its lower solubility, and thus greater sensitivity to
solubility is shown with data from Valtz et al. where solubility of error, than carbon dioxide. Improvements in the model can be
CO2 in pure water at 288.26 K and 1.941 MPa is measured to be made as more information on semi-clathrate crystal structure’s
0.01434 [47]. The closest data point in Table 6 is at 289.2 K and dependence on different conditions and vapor–liquid equilibrium
1.957 MPa with 40 wt% TBAB loading where the liquid mole data for TBAB–H2O–CH4 systems become available. The current
fraction is 0.013411. These solubility results at hydrate forming study is the first known to report the solubility of guest gas
conditions can be contrasted with another study showing that the compounds in semi-clathrate systems and presents the first model
presence of 6.7 wt% TBAB led to a four-fold increase in the explicitly focusing on predicting gas solubility values. Both the data
solubility of methane at atmospheric pressure and 25  C [48]. and model present a basis for future research into hydrate former
These studies demonstrate that any effect TBAB has on hydrate solubility in semi-clathrate systems.
former solubility at lower pressures and higher temperatures is
not comparable to its effect under hydrate forming conditions. Acknowledgements
Improvements to the two phase flash could be made using
vapor–liquid equilibrium data existing in literature for The authors are grateful to the Natural Sciences and Engineering
TBAB–H2O–CO2 systems[13]. However no such data was found Research Council of Canada (NSERC), McGill University and
for TBAB–H2O–CH4 systems, presenting problems since the specifically the McGill Engineering Doctoral Award (MEDA), les
methane system has the larger error. Another important Fonds Québécois de la Recherche sur la Nature et les Technologies
assumption is taking the hydration number as a constant for (FQRNT) and the Canada Research Chair Program (CRC) for financial
each guest gas. As discussed in the modeling section, the funding and support.
hydration number can change with pressure, temperature, TBAB
weight fraction and guest gases. Changing the hydration number Appendix A. Supplementary data
to be a function of TBAB weight fraction and guest gas type could
have a significant effect. Models can be adjusted as more Supplementary data associated with this article can be found, in
literature becomes available on the effect of these parameters the online version, at http://dx.doi.org/10.1016/j.fluid.2014.12.045.
on hydration number and structure. The assumption based on the
empty cage molar volume could be another source of error. References
Unfortunately, molar volumes of empty cages or thermal
expansion of TBAB hydrates have not been determined experi- [1] E.D. Sloan, C.A. Koh, Clathrate Hydrates of Natural Gases, third ed., CRC Press,
Boca Raton, FL, 2008.
mentally. However, a sensitivity analysis (10%) was done on the
[2] E.G. Hammerschmidt, Formation of gas hydrates in natural gas transmission
molar volume used and showed it does not have a significant lines, Ind. Eng. Chem. 26 (1934) 851–855.
effect on the final results of the model. The results of the [3] A. Eslamimanesh, A.H. Mohammadi, D. Richon, P. Naidoo, D. Ramjugernath,
sensitivity analysis are shown in Table 8. Improvements in the Application of gas hydrate formation in separation processes: a review of
experimental studies, J. Chem. Thermodyn. 46 (2012) 62–71.
model could also be achieved by including the hydrate former gas [4] M. Ricaurte, C. Dicharry, D. Broseta, X. Renaud, J.P. Torré, CO2 removal from a
in the liquid activity model (eNRTL). Despite these assumptions, CO2–CH4 gas mixture by clathrate hydrate formation using THF and SDS as
this model represents a good starting point for further water-soluble hydrate promoters, Ind. Eng. Chem. Res. 52 (2013) 899–910.
[5] A. Kumar, T. Sakpal, P. Linga, R. Kumar, Influence of contact medium and
semi-clathrate liquid phase modeling and can be improved as surfactants on carbon dioxide clathrate hydrate kinetics, Fuel 105 (2013)
more thermodynamic data on semi-clathrates becomes available. 664–671.
[6] J.P. Torre, M. Ricaurte, C. Dicharry, D. Broseta, CO2 enclathration in the presence
of water-soluble hydrate promoters: hydrate phase equilibria and kinetic
5. Conclusion studies in quiescent conditions, Chem. Eng. Sci. 82 (2012) 1–13.
[7] C.F.D.S. Lirio, F.L.P. Pessoa, A.M.C. Uller, Storage capacity of carbon dioxide
Equilibrium measurements of three phase hydrate–liquid– hydrates in the presence of sodium dodecyl sulfate (SDS) and tetrahydrofuran
(THF), Chem. Eng. Sci. 96 (2013) 118–123.
vapor systems containing tetra-n-butylammonium bromide
[8] D.L. Zhong, K. Ding, J. Yan, C. Yang, D.J. Sun, Influence of cyclopentane and SDS
(TBAB), water and either carbon dioxide or methane were on methane separation from coal mine gas by hydrate crystallization, Energy
undertaken. Pressure, temperature and liquid phase composition Fuels 27 (2013) 7252–7258.
[9] M. Arjmandi, A. Chapoy, B. Tohidi, Equilibrium data of hydrogen, methane,
measurements were performed and are consistent with other data
nitrogen, carbon dioxide, and natural gas in semi-clathrate hydrates of
in literature where available. A thermodynamic model was tetrabutyl ammonium bromide, J. Chem. Eng. Data 52 (2007) 2153–2158.
constructed by adapting common existing models to semi- [10] Z.G. Sun, L. Sun, Equilibrium conditions of semi-clathrate hydrate dissociation
clathrate systems. Equilibrium conditions were successfully for methane + tetra-n-butyl ammonium bromide, J. Chem. Eng. Data 55 (2010)
3538–3541.
modeled for TBAB–H2O–CO2 and TBAB–H2O–CH4 semi-clathrate [11] N. Mayoufi, D. Dalmazzone, W. Fürst, A. Delahaye, L. Fournaison, CO2
systems in the temperature range of 281–294 K, the pressure range enclathration in hydrates of peralkyl-(ammonium/phosphonium) salts:
168 J. Verrett et al. / Fluid Phase Equilibria 388 (2015) 160–168

stability conditions and dissociation enthalpies, J. Chem. Eng. Data 55 (2010) [30] C.-C. Chen, H.I. Britt, J.F. Boston, L.B. Evans, Local composition model for excess
1271–1275. Gibbs energy of electrolyte systems. Part I: single solvent single completely
[12] Z. Liao, X. Guo, Y. Zhao, Y. Wang, Q. Sun, A. Liu, C. Sun, G. Chen, Experimental dissociated electrolyte systems, AIChE J. 28 (1982) 588–596.
and modeling study on phase equilibria of semiclathrate hydrates of [31] H. Renon, J.M. Prausnitz, Local compositions in thermodynamic excess
tetra-n-butyl ammonium bromide + CH4, CO2, N2, or gas mixtures, Ind. Eng. functions for liquid mixtures, AIChE J. 14 (1968) 135–144.
Chem. Res. 52 (2013) 18440–18446. [32] L.S. Belveze, J.F. Brennecke, M.A. Stadtherr, Modeling of activity coefficients of
[13] W. Lin, D. Dalmazzone, W. Fürst, A. Delahaye, L. Fournaison, P. Clain, aqueous solutions of quaternary ammonium salts with the electrolyte–NRTL
Thermodynamic studies of CO2 + TBAB + water system: experimental equation, Ind. Eng. Chem. Res. 43 (2004) 815–825.
measurements and correlations, J. Chem. Eng. Data 58 (2013) 2233–2239. [33] J.S. Tse, Thermal expansion of the clathrate hydrates of ethylene oxide and
[14] M. Kwaterski, J.M. Herri, Thermodynamic modelling of gas semi-clathrate tetrahydrofuran, J. Phys. Colloques 48 (1987) C1-543–C1-549.
hydrate using the electrolyte NRTL model, 7th International Conference on Gas [34] R.H. Perry, D.W. Green, Perry’s Chemical Engineers’ Handbook, McGraw-Hill,
Hydrates, Edinburgh, United Kingdom, 2011. New York, 2008.
[15] A. Joshi, P. Mekala, J.S. Sangwai, Modeling phase equilibria of semiclathrate [35] Nist Janaf Thermochemical Tables, in: M.W. Jr. Chase, C.A. Davies, J.R. Jr.
hydrates of CH4, CO2 and N2 in aqueous solution of tetra-n-butyl ammonium Downey, D.J. Frurip, R.A. McDonald, A.N. Syverud (Eds.), Standard Reference
bromide, J. Nat. Gas Chem. 21 (2012) 459–465. Data Program, National Institute of Standards and Technology, Gaithersburg,
[16] A. Eslamimanesh, A.H. Mohammadi, D. Richon, Thermodynamic modeling 1985, pp. 20899.
of phase equilibria of semi-clathrate hydrates of CO2, CH4, or N2 + tetra-n- [36] O. Söhnel, P. Novotný, Densities of Aqueous Solutions of Inorganic Substances,
butylammonium bromide aqueous solution, Chem. Eng. Sci. 81 (2012) Elsevier, 1985.
319–328. [37] V. Belandria, A.H. Mohammadi, D. Richon, Volumetric properties of the
[17] P. Paricaud, Modeling the dissociation conditions of salt hydrates and gas (tetrahydrofuran + water) and (tetra-n-butyl ammonium bromide + water)
semiclathrate hydrates: application to lithium bromide, hydrogen iodide, and systems: experimental measurements and correlations, J. Chem. Thermodyn.
tetra-n-butylammonium bromide + carbon dioxide systems, J. Phys. Chem. B 41 (2009) 1382–1386.
115 (2011) 288–299. [38] H. Oyama, W. Shimada, T. Ebinuma, Y. Kamata, S. Takeya, T. Uchida, J. Nagao, H.
[18] S. Bergeron, P. Servio, CO2 and CH4 mole fraction measurements during Narita, Phase diagram, latent heat, and specific heat of TBAB semiclathrate
hydrate growth in a semi-batch stirred tank reactor and its significance to hydrate crystals, Fluid Phase Equilib. 234 (2005) 131–135.
kinetic modeling, Fluid Phase Equilib. 276 (2009) 150–155. [39] J. Lipkowski, V.Y. Komarov, T.V. Rodionova, Y.A. Dyadin, L.S. Aladko, The
[19] J.-S. Renault-Crispo, F. Lang, P. Servio, The importance of liquid phase structure of tetrabutylammonium bromide hydrate (C4H9)4NBr21/3H2O, J.
compositions in gas hydrate modeling: carbon dioxide–methane–water case Supramol. Chem. 2 (2002) 435–439.
study, J. Chem. Thermodyn. 68 (2014) 153–160. [40] L.S. Aladko, Y.A. Dyadin, T.V. Rodionova, I.S. Terekhova, Clathrate hydrates of
[20] H. Bruusgaard, J.G. Beltran, P. Servio, Solubility measurements for the tetrabutylammonium and tetraisoamylammonium halides, J. Struct. Chem. 43
CH4 + CO2 + H2O system under hydrate–liquid–vapor equilibrium, Fluid (2002) 990–994.
Phase Equilib. 296 (2010) 106–109. [41] W. Shimada, M. Shiro, H. Kondo, S. Takeya, H. Oyama, T. Ebinuma, H. Marita,
[21] A.H. Mohammadi, A. Eslamimanesh, V. Belandria, D. Richon, Phase equilibria Tetra-n-butylammonium bromide–water (1/38), Acta Crystallogr. C: Cryst.
of semiclathrate hydrates of CO2, N2, CH4, or H2 + tetra-n-butylammonium Struct. Commun. 61 (2005) o65–o66.
bromide aqueous solution, J. Chem. Eng. Data 56 (2011) 3855–3865. [42] J. Gholinezhad, C.A.B. Tohidi, Thermodynamic stability and self-preservation
[22] D.-L. Li, J.-W. Du, S.-S. Fan, D.-Q. Liang, X.-S. Li, N.-S. Huang, Clathrate properties of semi-clathrates in the methane + tetra-n-butyl ammonium
dissociation conditions for methane + tetra-n-butyl ammonium bromide bromide + water system, 7th International Conference on Gas Hydrates,
(TBAB) + water, J. Chem. Eng. Data 52 (2007) 1916–1918. Edinburgh, United Kingdom, 2011.
[23] N. Ye, P. Zhang, Equilibrium data and morphology of tetra-n-butyl ammonium [43] P. Servio, P. Englezos, Measurement of dissolved methane in water in
bromide semiclathrate hydrate with carbon dioxide, J. Chem. Eng. Data 57 equilibrium with its hydrate, J. Chem. Eng. Data 47 (2002) 87–90.
(2012) 1557–1562. [44] Thermodynamics Research Center, NIST Boulder Laboratoriess, M. Frenkel,
[24] J. Deschamps, D. Dalmazzone, Dissociation enthalpies and phase equilibrium director, Thermodynamics Source Database, NIST Chemistry WebBook, NIST
for TBAB semi-clathrate hydrates of N2, CO2, N2 + CO2 and CH4 + CO2, J. Therm. Standard Reference Database Number 69. Linstrom, P.J., Mallard, W.J. (Eds.),
Anal. Calorim. 98 (2009) 113–118. National Institute of Standards and Technology, National Institute of Standards
[25] M.A. Trebble, P.R. Bishnoi, Extension of the Trebble–Bishnoi equation of state and Technology, Gaithersburg MD, 20899, http://webbook.nist.gov.
to fluid mixtures, Fluid Phase Equilib. 40 (1988) 1–21. [45] D.C. Montgomery, G.C. Runger, Applied Statistics and Probability for Engineers,
[26] M.A. Trebble, P.R. Bishnoi, Development of a new four-parameter cubic fourth ed., Wiley, Hoboken, NJ, 2007.
equation of state, Fluid Phase Equilib. 35 (1987) 1–18. [46] K. Lekvam, P.R. Bishnoi, Dissolution of methane in water at low temperatures
[27] W.R. Parrish, J.M. Prausnitz, Dissociation pressures of gas hydrates formed by and intermediate pressures, Fluid Phase Equilib. 131 (1997) 297–309.
gas mixtures, Ind. Eng. Chem. Proc. Des. Dev. 11 (1972) 26–35. [47] A. Valtz, A. Chapoy, C. Coquelet, P. Paricaud, D. Richon, Vapour–liquid equilibria
[28] J.H. van der Waals, J.C. Platteeuw, Clathrate solutions, Adv. Chem. Phys. 2 in the carbon dioxide–water system, measurement and modelling from
(1959) 1–55. 278.2 to 318.2 K, Fluid Phase Equilib. 226 (2004) 333–344.
[29] J.B. Klauda, S.I. Sandler, A fugacity model for gas hydrate phase equilibria, Ind. [48] A. Feillolay, M. Lucas, Solubility of helium and methane in aqueous
Eng. Chem. Res. 39 (2000) 3377–3386. tetrabutylammonium bromide solutions at 25 and 35 , J. Phys. Chem. 76
(1972) 3068–3072.

You might also like