Sunteți pe pagina 1din 11

VOL.

45
NOS.
4
A P R I L
2 0 1 2
J OURNAL OF
CHEMI CAL ENGI NEERI NG
OF J APAN
The Society of
Chemical Engineers,
Japan
JCEJAQ 45(4)
245-310(2012)
ISSN 0021-9592
[ VOL. 45, NO. 4, APRIL 2012]
Process Systems Engineering and Safety
Development of Environment for Logical Process Safety Management Based on the Business Process
Model (Journal Review)
Yukiyasu Shimada 245
Transport Phenomena and Fluid Engineering
Operation for Fine Particle Dispersion in Shear-Thinning Fluid in a Stirred Vessel
Takafumi Horie, Nobuhiro Sakano, Eiji Aizawa, Norihisa Kumagai and Naoto Ohmura 258
Self-Organization of Polymer Films Using Single and Mixed Solvents on Chemically Patterned Surfaces
Shohei Yasumatsu, Hirotaka Ishizuka, Koichi Nakaso, Patience Oluwatosin Babatunde and Jun Fukai 265
Particle Engineering
Acoustic Efect on Induction of Cerium Carbonate in Reaction Crystallization
Xin-Kuai He, Rayoung Kim, Dongmin Shin and Woo-Sik Kim 272
Denition of the New Mean Particle Size Based on the Settling Velocity in Liquid (SC)
Rondang Tambun, Masamitsu Shimadzu, Yuichi Ohira and Eiji Obata 279
Separation Engineering
Improvement of the Deethanizing and Depropanizing Fractionation Steps in NGL Recovery Process Using
Dividing Wall Column
Nguyen Van Duc Long and Moonyong Lee 285
Chemical Reaction Engineering
Selective Hydrogenation of Unsaturated Aldehyde in a GasLiquidLiquidSolid Batch Reactor (SC)
Hiroshi Yamada, Takashi Miura, Tomohiko Tagawa and Shigeo Goto 295
Process Systems Engineering and Safety
Nonlinear Process Control Using a Direct Adaptive PI Controller
Chyi-Tsong Chen and Shih-Tien Peng 298
Materials Engineering and Interfacial Phenomena
Production of Fine Organic Crystalline Particles by Using Milli Segmented Flow Crystallizer
Shoji Kudo and Hiroshi Takiyama 305
This article appeared in the Journal of Chemical Engineering of Japan.
The attached copy is provided to the author for non-commercial research, education use and sharing with colleagues.
Other uses listed below are prohibited:
- Reproduction,
- Commercial use,
- Posting to personal, institutional or third party websites.
Vol. 45 No. 4 2012 285 Copyright 2012 The Society of Chemical Engineers, Japan
Journal of Chemical Engineering of Japan, Vol. 45, No. 4, pp. 285294, 2012
Improvement of the Deethanizing and Depropanizing
Fractionation Steps in NGL Recovery Process Using
Dividing Wall Column
Nguyen Van Duc L and Moonyong L
chool of Chemical Engineering, Yeungnam University, Gyeongsan 712-749, South Korea
Keywords: Distillation, Dividing Wall Column, DWC, Natural Gas Liquid, Fully Thermally Coupled Distillation Column,
Heat Integration
In this work, our aim is to utilize a dividing wall column to improve the performance of the deethanizing and depro-
panizing fractionation steps in natural gas liquid processing. Starting from an initial conventional column sequence,
the initial designs of the conventional dividing wall column and a top dividing wall column are obtained by maintaining
the number of trays. In succession, they are optimized to reduce the energy consumption using factorial design. The
results show that the conventional dividing wall column and the top dividing wall column can ofer many benefts to the
system, e.g., curbing the operating cost including refrigeration cost, and minimizing the reboiler and condenser duty.
Furthermore, by using a dividing wall column, both the purity of ethane and its recovery rate are increased. The infu-
ence of utility prices on the operating cost saving of the conventional and the top dividing wall columns is also inves-
tigated. In addition, to further enhance the dividing wall column performance, heating is integrated on the top and an
interreboiling system is installed at the bottom section of the dividing wall column.
Introduction
Te natural gas (NG) industry, which frst dates back to
the early 1900s in the United States, is still an evolving feld.
Although the primary use of NG is as a fuel, it can also
be a source of hydrocarbons for petrochemical feed stocks
and a major source of elemental sulfur, which is an impor-
tant industrial chemical (Kidnay and Parrish, 2006). Due
to the clean burning characteristics and the ability to meet
stringent environmental requirements, the demand for nat-
ural gas has increased considerably over the past few years.
Terefore, the recovery of natural gas liquids (NGL) has be-
come more and more economically attractive as a number of
components in natural gas are ofen isolated from the bulk
gas and then sold separately. Consequently, there are numer-
ous process confgurations and methods known to increase
NGL recovery from a feed gas. In addition, these processes
can provide potential enhancements to the overall facility
availability and project economics when employing the in-
tegrated concept (Elliot et al., 2005; Mak, 2006). However,
one of the challenges in the integration of an NGL recovery
process is the diference in column operating pressures.
Distillation is the primary separation process used in
chemical processing industries. While this unit operation
has many advantages, one major drawback is its large energy
requirement (Schultz et al., 2002). Tis can signifcantly in-
fuence the overall plant proftability. Terefore, the increas-
ing cost of energy has forced industry to reduce its energy
consumption. In addition, the tendency to rationalize and to
reduce the usage of fossil energy resources, and the tighter
environmental regulations caused by, for example, the efort
to prevent climate change, have generated the need to em-
ploy new and more efcient separation methods (Knapp and
Doherty, 1990; Malinen and Tanskanen, 2009).
Several approaches have been proposed to develop alter-
natives for energy savings. One of them is the integration
of the column with the rest of the process (Linnhof et al.,
1983; Jimnez et al., 2003). Figure 1(a) illustrates the di-
rect conventional distillation sequence for ternary distilla-
tion. Tere is a reboiler and a condenser in every distillation
column. Te number of product fows in the columns is
always the same, i.e. two. However, in a more novel distilla-
tion scheme, i.e. in a prefractionator scheme, as illustrated
in Figure 1(b), there is one extra product fow, known as a
side stream, in the second column of the column sequence
(Malinen and Tanskanen, 2009). Trough the implementa-
tion of two interconnecting streams (one in the vapor phase
and the other in the liquid phase) between the two columns,
thermally coupled distillation systems can be built; such sys-
tems have been particularly studied for the separation of ter-
nary mixtures, and have been shown to provide signifcant
energy savings with respect to both the conventional direct
and indirect distillation sequences. Te three most widely
analyzed thermally coupled systems are the sequence with
a side rectifer, the sequence with a side stripper and the
fully thermally coupled system (or Petlyuk column), which
are illustrated in Figures 2(a), (b) and Figure 3(a), respec-
tively (Glinos et al., 1986; Fidkowski and Krolikowski, 1990;
Triantafyllou and Smith, 1992; Hernndez and Jimnez,
1996; Jimnez et al., 2003).
Research Paper
Received on August 30, 2011; accepted on December 3, 2011
Correspondence concerning this article should be addressed to M. Lee
(E-mail address: mynlee@yu.ac.kr).
286 Journal of Chemical Engineering of Japan
Of the above possible thermal coupling arrangements, the
side stripper and the side rectifer have been widely used
in refnery distillation and cryogenic air separation, re-
spectively (Seidel, 1935; Watkins, 1979; Amminudin et al.,
2001). Te Petlyuk column provides a fully interconnected
structure, with two thermal couplings that result in vapor
interconnections fowing back and forth between the col-
umns (Hernndez et al., 2005). Tese sequences have been
shown to provide energy savings of up to 30% over the con-
ventional direct and indirect distillation sequences (Tedder
and Rudd, 1978; Alatiqi and Luyben, 1985; Triantafyllou
and Smith, 1992; Finn, 1993; Hernndez and Jimnez, 1996,
1999; Blancarte-Palacios et al., 2003).
However, the Petlyuk column undergoes strong interac-
tions between the two columns because of their thermal in-
tegration, which causes difculties in design and operation.
To solve this problem as well as to reduce the capital cost, a
vertical wall is installed in the central section of the column,
dividing the column into the prefractionator and the main
column, as schematically drawn in Figure 3(b). Tis ar-
rangement is called the dividing wall column (DWC), which
is conceptually the same as the Petlyuk column due to their
thermodynamically equivalent arrangements (Amminudin
et al., 2001). By moving the dividing wall to the top or
the bottom of the column, the top dividing wall column
(TDWC) or the bottom dividing wall column (BDWC) can
be constituted, respectively.
Although DWCs can potentially reduce energy use and
investment costs (Midori et al., 2000; Oluji et al., 2009;
Dejanovi et al., 2010; Long et al., 2010; Long and Lee,
2011), their design and optimization is challenging, involv-
ing the solving of multivariable problems with variables that
interact with each other and need to be optimized simulta-
neously. A common difculty associated with detailed simu-
lation is related to the estimation of the number of stages in
each section (Dejanovi et al., 2010). Statistical approaches
have shown potential for this, with DWCs' structures hav-
ing been optimized by response surface methodology (Long
and Lee, 2012). Tis method can be easily and efciently
implemented using Hysys and Minitab. Another statistical
approach, factorial design, is widely used in experimen-
tal investigation (Montgomery, 1991; Cestari et al., 1996;
Persson and O. strm, 1997; Carvalho et al., 1999; Cestari
et al., 2008). However, factorial design can only give relative
values. Regardless, this method is a useful method to design
experiments or simulation in both laboratory and industrial
settings.
A set of improved deethanizing and depropanizing frac-
tionation steps in NGL processing, such as interreboiling,
intercondensing, thermal coupling, and thermo-mechan-
ical coupling, has been studied (Manley, 1997). Inspired
by this study, the our aim of study is, therefore, to further
enhance the performance of the deethanizing and depro-
panizing fractionation steps in NGL processing using DWC
and heat integration. Factorial design was employed to fnd
the optimum structure of DWC. Furthermore, a comparison
(a)
(b)
Fig. 1 Schematic diagram of the (a) direct sequence and (b) prefrac-
tionator arrangement
(a)
(b)
Fig. 2 Schematic diagram of the thermally coupled distillation se-
quences: (a) system with side rectifer and (b) system with side
stripper
Vol. 45 No. 4 2012 287
between the conventional DWC and the TDWC was per-
formed.
1. Conventional Distillation Sequence
Figure 4 illustrates the conventional distillation sequence
and its current operating conditions. Te critical tempera-
ture for ethane is approximately 32.3C; therefore, the NGL
deethanizer column requires a refrigerated condenser when
producing relatively pure ethane. In order to minimize the
refrigeration costs, the deethanizer was designed and oper-
ated at relatively high pressures of about 450 psia (31.0 bar).
Te resulting condenser temperature of nearly 13.3C (for
relative pure ethane) is also high enough to preclude the
formation of hydrates which can plug the equipment. With
18 theoretical stages, the deethanizer column produces a
partially condensed overhead vapor stream. Whereas, the
depropanizer, possessing 34 theoretical stages, was de-
signed and operated at about 250 psia (17.2 bar), as com-
mercial propane can be condensed with cooling water at
this pressure (Manley, 1997). Te feed composition, tem-
perature, and pressure are described in Table 1. Te simula-
tion work was performed using the simulator Aspen Hysys
V7.1 (Aspen Technology, Inc.). Te PengRobinson equa-
(a)
(b)
Fig. 3 Schematic diagram of the (a) Petlyuk column and (b) dividing
wall column
Fig. 4 Simplifed fow sheet illustrating the separation train of two
conventional columns
Table 1 Feed conditions of the mixture
Feed conditions
Component Mass fow [kg/h]
Liquid volume
fractions [%]
Methane 991.66 0.50
Ethane 87179.69 37.00
Propane 87268.17 26.00
i-Butane 26803.59 7.20
n-Butane 57180.35 14.80
i-Pentane 20649.83 5.00
n-Pentane 14600.65 3.50
n-Hexane 17559.16 4.00
n-Heptane 9099.56 2.00
Temperature [C] 55.83
Pressure [bar] 31.37
Table 2 Column hydraulics, energy performance, and product speci-
fcations of the existing columns sequence
Deethanizer Depropanizer
Number of trays 18 34
Tray type Sieve Sieve
Column diameter [m] 6.2 4.9
Number of fow paths 1 1
Tray spacing [mm] 609.6 609.6
Max fooding [%] 85.99 83.51
Condenser duty [MW] 14.67 26.03
Reboiler duty [MW] 28.95 21.54
Purity of C
2
[Liq. Vol.
Fractions %]
95.13
Recovery of C
2
[%] 94.58
Purity of C
3
[Liq. Vol.
Fractions %]
90.14
Recovery of C
3
[%] 92.47
288 Journal of Chemical Engineering of Japan
tion of state that supports the widest range of operating con-
ditions and the greatest variety of systems was used for the
prediction of vaporliquid equilibrium of these simulations
(Aspen Technology, 2009). Table 2 presents the conditions
and product specifcations for each column in the conven-
tional distillation sequence. From the base case simulation
model, it shows that the energy consumption of the deetha-
nizer and depropanizer are 28.95 and 21.54 MW, respec-
tively.
All columns were designed with the maximum fooding
of near 85%, so as to prevent fooding in the columns. To
determine the maximum fooding of a particular column,
the rating mode is simulated with column internal specif-
cations including the type of trays, column diameter, tray
spacing, and number of passes. Table 2 lists the parameters
necessary to defne the existing hydraulic features of the col-
umns.
It should be noted that, in order to estimate the operating
cost of the proposed sequences, the utility cost data from
published literatures (Krajnc and Glavic, 1996; Biegler et al.,
1997), shown in Table 3, was used. If a process stream needs
to be operated below 27C, cooling is certainly required,
and thus a refrigeration cycle needs to be considered.
Refrigerating a stream is far more expensive (on an energy
basis) than lowering the temperature with cooling water, or
even raising the temperature with steam. Consequently, re-
frigeration is generally not a desirable option and other al-
ternative processes should be considered frst (Biegler et al.,
1997). Importantly, in the conventional column sequence,
the condensers in the deethanizer and depropanizer utilize a
refrigeration system and water, respectively.
2. Integrating Using CDWC
2.1 Conventional dividing wall column (CDWC)
In the NGL recovery process, the operating pressure is
an important factor afecting the processing cost. Increasing
the column pressure increases the temperature of the top
vapor stream, i.e. decreases the refrigeration cost. However,
it increases the burden on the reboiler, the condenser and
also the vessel wall, resulting in a thicker column with a
relatively large diameter. Terefore, the optimum pressure of
the column is achieved by balancing the investment cost and
operating cost. Furthermore, it has been discovered that the
inefcient remixing of propane and butane in the bottom of
the deethanizer can be eliminated by reducing the deetha-
nizer pressure to match the depropanizer and by thermally
coupling the two columns (Manley, 1997). It is also known
that a relatively low pressure (about 17.5 bar) deethanizer
requires more refrigeration utilities, but has a much smaller
diameter and requires less reboiler duty than a high pressure
(about 31.0 bar) one. In addition, the wall thickness, weight
and cost of the column will be further reduced because of
the lower operating pressure. Terefore, the DWC, operated
at approximately 17.5 bar, is applied herein to minimize the
investment cost and the energy consumption.
A shortcut design procedure was applied to construct the
initial CDWC structure (Amminudin et al., 2001; Long et
al., 2010; Lee et al., 2011; Long and Lee, 2011) based on
the conventional column confguration shown in Figure 5.
In this conventional confguration, the frst column corre-
sponds to the prefractionator section in the CDWC. Te
rectifying section of the second column and the stripping
section of the third column represent the top and bottom
sections of the CDWC, respectively. Both the stripping sec-
tion of the second column and the rectifying section of the
third column are equivalent to the divided wall section of
the CDWC. Furthermore, both the bottom stream from the
second column and the top stream from the third column
refer to the side stream of the CDWC (Long et al., 2010).
Consequently, the structure of the CDWC can be divided
into four sections: the prefractionator section for the feed
mixture; the top and bottom sections above and below the
divided wall; and the divided wall section.
Te new CDWC was maintained with 34 stages for a fair
comparison with the other sequences. Te main design vari-
ables including the internal vapor and liquid fow to the
prefractionator, the number of trays in the top (N1), middle
(N2), feed stripping (N3), and side product (N4) section
were optimized. In this work, the number of trays in the
Table 3 Utilities cost data
Utility Price [USD/(kW a)]
Refrigerant 377
Cooling water 90
Steam 300
Electricity 104
Fig. 5 A three-column distillation system for the initial design of
DWC structure
Vol. 45 No. 4 2012 289
dividing wall sections both for the prefractionator and the
main fractionators are fxed as the same values as N2. A fac-
torial design was employed to analyze their interactive efect
and to reach an optimum in terms of the reboiler consump-
tion. Te simulation run data were ftted to a second order
polynomial model, and subsequently regression coefcients
were obtained. Te generalized second order polynomial
model used in the factorial design is as given by Eq. (1).
= <
= + + +
0
1
k
i i ij i j
i i j
Y X X X


(1)
Here, Y is the predicted response (energy consumption), X
i

is the uncoded or coded value of the variables,
0
is a con-
stant,
i
is the main efect coefcient,
ij
is the interaction
efect coefcient, and is the error. Te second order poly-
nomial coefcients and optimum value of the reboiler con-
sumption were obtained using the Minitab sofware.
Table 4 lists the factors and levels used in this case study.
A total of 16 simulation runs were used to fnd the opti-
mum values of the 4 parameters associated with the CDWC
structure. For each run, the internal vapor and liquid fow to
the prefractionator were varied to minimize the energy con-
sumption while meeting the required product purity. Te
factorial design can cover the main and interaction efects of
the parameters over the whole range of selected parameters.
According to the sparsity-of efects principle in factorial de-
sign, it is most likely that the main (single factor) efects
and two-factor interactions are the most signifcant efects,
and the higher order interactions are negligible (Nejad et
al., 2011). In other words, higher order interactions such
as three-factor interactions are very rare and considered as
the residual, which are dispersed randomly. Figure 6 pres-
ents the main efect plots of design variables N1, N2, N3
and N4 on the reboiler duty. Te analysis of the efect of
the principal factors illustrated that, in the considered range
of parameters, the side product section (N4) is the most
signifcant variable in achieving the minimum energy con-
sumption. Figure 7 shows the plots of interactions between
the N1, N2, N3 and N4. Te best result of the reboiler con-
sumption was found at the coded levels with the number
of trays in the top, middle, feed stripping, and side product
section of 1, 1, 1, and 1, respectively. From these re-
sults, the natural value of the variables can be derived. Ten,
the recycle vapor and liquid fow were optimized to achieve
the minimum energy consumption and a satisfactory level
of product purity. Figure 8 reveals a simplifed fow sheet
illustrating the resulting CDWC system. According to the
rigorous simulation results, the duty of the condenser and
reboiler are 18.92 and 28.60 MW, respectively. Using the
CDWC can save 31.74% of the operating cost. Furthermore,
compared with the two columns with diameters of 6.2 and
Fig. 6 Main efects plots of design variables N1, N2, N3, and N4 to
the reboiler duty
Fig. 7 Interaction plots between main design variables N1, N2, N3,
and N4
Table 4 Coded levels of factors
Factor
Levels
1 1
Top section 2 4
Middle section 18 22
Feed stripping section 11 15
Side product section 3 6
Fig. 8 Simplifed fow sheet illustrating the CDWC
290 Journal of Chemical Engineering of Japan
4.9 m, respectively, the use of CDWC with a diameter of
5.5 m, as shown in Table 5, could give a saving of 33.63%
in the total annual cost (TAC) when operating at low pres-
sure. Furthermore, the purity of C
2
is increased from 95.13
to 97.73% and its recovery also is improved from 94.58 to
98.21% as compared to the conventional column sequence.
2.2 Heat integrated CDWC
In the present study, the top vapor stream is divided fur-
ther into two gas streams. Te frst stream, a refux stream,
is compressed to a pressure greater than that of the CDWC.
Tis refux is then cooled in a heat exchanger due to the
second stream, which then is water-cooled in a condenser.
Figure 9 is a schematic illustration of this confguration,
particularly showing how a portion of the overhead vapor
stream from the deethanizer is recycled back to the top of
the deethanizer. With this confguration, the temperatures of
the top vapor stream of the CDWC afer being compressed
and the refux stream are approximately 62 and 32C, re-
spectively, thereby allowing for water to condense both
streams.
To minimize the energy consumption, the bottom liquid
product can be subcooled while interreboiling the bottom
section of the DWC (Manley, 1997). As a result, this sys-
tem achieves a substantial saving of 47.58% in the operat-
ing cost as compared to the conventional column sequence.
Note that the conventional column sequence with the same
integration strategy can save up 18.28% in terms of operat-
ing cost.
3. Integrating Using TDWC
3.1 TDWC
Te thermally coupled low pressure deethanizer and
depropanizer were, therefore, studied to reduce the energy
consumption (Manley, 1997). However, this process requires
two columns, which causes an increase in the investment
cost. Tus, to decrease the refrigeration cost and investment
cost, the TDWC, which is equivalent to a combination of
both columns, was applied. Consequently, the system shown
in Figure 10 is built by moving the wall up to the top of
the column, thereby separating this column into three sec-
tions: one is the prefractionator section for the feeding the
mixture with its own condenser (C1) to produce C
2
, another
is the bottom section below the divided wall, and the last
is the divided wall section. Compared to the CDWC, the
TDWC does not have the middle section and has one more
condenser. Tis could result in a higher investment cost as
compared to the CDWC. Furthermore, the liquid split ratio
in this column cannot be adjusted except for the vapor split
ratio, i.e. this kind of column could not demonstrate the
full advantage of the conventional DWC in terms of energy
saving.
Te main purpose in this part is to decrease the refrig-
eration duty afecting the operation of the natural gas pro-
cessing to a great extent. Tus, this column can be utilized
to integrate two columns, whose condensers are cooled by
diferent kinds of cooling resources; particularly, refrigera-
tion needs to be used in the deethanizer, while the depro-
Fig. 9 Simplifed fow sheet illustrating the CDWC with modifca-
tions
Fig. 10 Simplifed fow sheet illustrating the TDWC
Table 5 Design results for the CDWC
Parameters Value
Liquid split* 0.36
Vapor split* 0.65
Refux rate [kg/h] 194706
Reboiler duty [MW] 28.60
Diameter [m] 5.50
Top pressure [bar] 17.50
Max fooding [%] 86.93
Purity of C
2
[Liq. Vol. Fractions %] 97.73
Recovery of C
2
[%] 98.21
Purity of C
3
[Liq. Vol. Fractions %] 90.14
Recovery of C
3
[%] 92.47
* Note that the split is defned as the ratio between the fow into the prefrac-
tionator to the total fow into the dividing wall section.
Vol. 45 No. 4 2012 291
panizer just needs water to condense the top vapor stream.
As a result, the refrigeration cost decreases; in particular,
as compared with the conventional column sequence, the
TDWC requires less than 0.79 MW. Furthermore, the con-
denser (C2) and the reboiler duty can be curbed by 46.36
and 25.73%, respectively. In addition, the purity and recov-
ery of C
2
also increase from 95.13 to 96.71% and 94.58 to
96.77%, respectively, as compared with the conventional col-
umn sequence. Te cross sectional area of the top section
of the TDWC is the sum of the prefractionator section area
and the top section area of the main fractionator. Te diam-
eter of the middle section can be calculated from the cross
sectional area of the middle section. Table 6 lists the diam-
eter of the top and bottom sections of the TDWC. Based on
the results in Table 6, the length of 6.1 m was chosen as the
diameter of TDWC. Overall, using TDWC can save TAC
by 24.02% as compared to the conventional distillation se-
quence.
In the literature, there is almost no data on the study of
TDWC or any comparison between the CDWC and TDWC.
Tus, a comparison of their performances was performed in
our study. As mentioned earlier, the TDWC has one more
condenser and the liquid split ratio cannot be adjusted as
compared to CDWC, causing a higher investment cost and
more steam consumption, respectively. Table 7 summarizes
the comparison of the performances between CDWC and
TDWC. It can be seen that the TDWC requires less refrig-
erator duty, having a cost that is afected by the electricity
utility price, and more reboiler duty, having a cost that is
mainly afected by the fuel price. Furthermore, the TDWC
requires a larger diameter than the CDWC. In addition, the
infuence of utility prices on the operating cost saving of the
CDWC and TDWC was studied with varying the ratio of re-
frigerant/steam cost. Table 8 reveals that when the cost ratio
of refrigerant/steam almost triples, the TDWC ofers more
benefts in terms of the operating cost saving as compared to
Table 6 Diameter of the TDWC
Prefractionator Mainfractionator (top) TDWC (top) TDWC (bottom) TDWC
Diameter [m] 4.5 3.1 5.5 6.1 6.1
Max fooding [%] 83.29 82.13 84.64
Position of section 17th tray34th tray 1st tray16th tray
Table 7 Comparison of diferent structural alternatives for improving the performance of deethanizer and depropanizer
Structural alternative
Conventional column sequence
CDWC TDWC
Conventional column sequence
with heat integration
CDWC with
heat integration
TDWC with
heat integration
Deethanizer Depropanizer Deethanizer Depropanizer
Tray number 18 34 34 34 18 34 34 34
Column diameter
[m]
6.2 4.9 5.5 6.1 6.2 4.9 5.5 6.1
Refrigeration duty
[MW]
14.67 0.00 18.92 13.88 0 0.00 0.00 0.00
Cooling-water
Condenser duty
[MW]
0.00 26.03 0.00 13.96 16.38 26.25 31.01 35.38
Compressor duty
[MW]
0.00 0.00 0.00 0.00 3.34 0.00 14.30 10.41
Reboiler duty [MW] 28.95 21.54 28.60 37.50 28.91 19.85 25.96 33.68
Reboiler duty saving
[%]
0.00 43.36 25.73 3.43 48.58 33.29
Operating cost sav-
ing [%]
0.00 31.74 22.94 18.28 47.58 37.57
TAC saving [%] 0.00 33.63 24.02 16.61 43.95 33.99
Purity of C
2
[Liq.
Vol. Fractions %]
95.13 97.73 96.71 95.13 97.72 97.18
Recovery of C
2
[%] 94.58 98.21 96.77 94.58 98.22 97.45
Table 8 Infuence of utility prices on the operating cost saving of the CDWC and TDWC
Price [USD/(kW a)] Refrigerant 300 377 600 800 1000
Steam 300 300 300 300 300
Cost ratio of refrigerant/steam 1.00 1.25 2.00 2.67 3.33
Operating cost saving [%] CDWC 34.88 31.74 24.19 18.85 14.49
TDWC 23.85 22.94 20.75 19.21 17.95
292 Journal of Chemical Engineering of Japan
the CDWC. Terefore, building a new plant with the CDWC
or TDWC has to be considered carefully based on the en-
ergy price and the equipment cost.
3.2 Heat integrated TDWC
In order to take further advantage of energy efciency,
the integrated confguration applied to the CDWC was also
employed to the TDWC, as shown in Figure 11. Table 7
shows that the operating cost and the TAC can be curbed
up to 37.57 and 33.99% as compared with the conventional
column sequence. Te results indicate that the CDWC with
heat integration could obtain the largest TAC saving as com-
pared to other sequences.
Conclusions
In this paper, the integration of the deethanizer and
depropanizer has been presented to reduce costs and en-
ergy consumption. Factorial design is employed to opti-
mize the DWC with an easy and efcient implementation
of Hysys and Minitab. Te interaction between the variables
can be identifed or quantifed by the factorial design. Te
use of CDWC can gain many benefts as well as decreasing
the TAC by 33.63%. When the refrigerant cost is too high,
the TDWC, combining the deethanizer and depropanizer
into a single fractionation column, reduces the pressure of
the deethanizer to that of the depropanizer and locates the
deethanizer and rectifying section of the depropanizer in
the top portion of a single dividing wall column, and can
be applied as a replacement. By using this kind of column,
not only the energy (refrigeration energy, reboiler and con-
denser duty) and the investment cost (replacing two col-
umns with a column and eliminating the deethanizer re-
boiler) can be curbed, but also the purity and recovery of
ethane can be increased. In addition, reducing the number
of columns, condensers, reboilers, as well as the vessel wall
thickness, and weight as a result of the lower operating pres-
sure, can also provide benefts in the investment cost. Based
on these results, a comparison of the performances between
the CDWC and TDWC has been analyzed. Furthermore, to
further enhance the DWC performance, heating has been
integrated on the top and an interreboiling system has been
installed at the bottom section of the DWC.
Acknowledgment
Tis research was supported by a grant from the Gas Plant R&D
Center funded by the Ministry of Land, Transportation and Maritime
Afairs (MLTM) of the Korean government.
Appendix: Column Cost Correlations
a. Sizing the column: Te column diameter was determined by the
column fooding condition that fxes the upper limit of vapor veloc-
ity. Operating velocity is normally between 70 and 90% of the fooding
velocity (Sinnott, 2005; Premkumar and Rangaiah, 2009); In this study,
85% of the fooding velocity was used.
b. Capital cost: Guthries modular method was applied (Biegler et
al., 1997). Te investment cost for conventional distillation is the total
cost of the column and the auxiliary equipment, such as the reboiler
and condenser. For the DWC, it included the additional dividing wall
cost. In this study, for cost updating, the Chemical Engineering Plant
Cost Index of 575.4 was used.
Updated bare module cost ( ) ( 1) BMC UF BC MPF MF = + (2)
where UF is the update factor:
present cost index / base cost index UF = (3)
BC is bare cost: For vessels:
0
0 0

L D
BC BC
L D



= (4)
For the heat exchanger:
0
0

S
BC BC
S



= (5)
Area of heat exchanger:

Q
S
U T
= (6)
where MPF is the material and pressure factor; MF is the module factor
(a typical value), which is afected by the base cost. D, L, and S represent
the diameter, length and area, respectively.
Compressor cost was interpolated (Peters and Timmerhous, 1991).
c. Operating cost (Op):
steam CW refrigerant electricity
Op C C C C = + + + (7)
where C
stream
is the cost of steam; C
CW
is the cost of cooling water;
C
refrigerant
is the cost of refrigerant and C
electricity
is the cost of electricity.
d. Total annual cost (TAC) (Smith, 1995)
(1 )
capital cost
(1 ) 1
n
n
i i
TAC Op
i
+
= +
+
(8)
where i is the fractional interest rate per year and n is the number of
years.
Nomenclature
B = bottom product fow [kg/h]
C
2
= ethane []
C
3
= propane []
C1 = the condenser of the prefractionator []
C2 = the condenser of the main fractionator []
D = top product fow [kg/h]
F = feed fow [kg/h]
R
L
= liquid split ratio []
R
V
= vapor split ratio []
Fig. 11 Simplifed fow sheet illustrating the TDWC with modifca-
tions
Vol. 45 No. 4 2012 293
S = side product fow [kg/h]
Subscript and Superscript
L = liquid
V = vapor
Literature Cited
Alatiqi, I. M. and W. L. Luyben; Alternative Distillation Confgura-
tions for Separating Ternary Mixtures with Small Concentration
of Intermediate in the Feed, Ind. Eng. Chem. Process Des. Dev., 24,
500506 (1985)
Amminudin, K. A., R. Smith, D. Y. C. Tong and G. P. Towler; Design
and Optimization of Fully Termally Coupled Distillation Col-
umns: Part 1: Preliminary Design and Optimization Methodol-
ogy, Chem. Eng. Res. Des., 79, 701715 (2001)
Aspen Technology ed.; Aspen HYSYS Termodynamics COM Inter-
face, Version V7.1, Aspen Technology Inc., Burlington, U.S.A.
(2009)
Biegler, L. T., I. E. Grossmann and A. W. Westerberg; Systematic Meth-
ods of Chemical Process Design, pp. 110141, Prentice Hall Inc.,
Upper Saddle River, U.S.A. (1997)
Blancarte-Palacios, J. L., M. N. Bautista-Valads, S. Hernndez Cas-
tro, V. Rico-Ramrez and A. Jimnez; Energy-Efcient Designs
of Termally Coupled Distillation Sequences for Four-Component
Mixtures, Ind. Eng. Chem. Res., 42, 51575164 (2003)
Carvalho, M. L., J. M. S. Cabral and M. R. Aires-Barros; Cutinase
Stability in AOT Reversed Micelles: System Optimization Using
the Factorial Design Methodology, Enzyme Microb. Technol., 24,
569576 (1999)
Cestari, A. R., C. Airoldi and R. E. Bruns; A Fractional Factorial De-
sign Applied to Organofunctionalized Silicas for Adsorption Opti-
mization, Colloids Surf., A, 117, 713 (1996)
Cestari, A. R., E. F. S. Vieira and J. A. Mota; Te Removal of an Anion-
ic Red Dye from Aqueous Solutions Using Chitosan BeadsTe
Role of Experimental Factors on Adsorption Using a Full Factorial
Design, J. Hazard. Mater., 160, 337343 (2008)
Dejanovi, I., L. Matijaevi and . Oluji; Dividing Wall Column-A
Breakthrough towards Sustainable Distilling, Chem. Eng. Process.,
49, 559580 (2010)
Elliot, D., W. R. Qualls, S. Huang and J. J. (Roger) Chen; Beneft of
Integrating NGL Extraction and LNG Liquefaction Technology,
AIChE Spring National Meeting, 5th Topical Conference on Natu-
ral Gas Utilization (TI) Session 16c-Gas, pp. 1014, Atlanta, U.S.A.
(2005)
Fidkowski, Z. and L. Krolikowski; Energy Requirements of Noncon-
ventional Distillation Systems, AIChE J., 36, 12751278 (1990)
Finn, A. J.; Consider Termally Coupled Distillation, Chem. Eng.
Prog., 10, 4145 (1993)
Glinos, K., I. Nikolaides and F. Malone; New Complex Column Ar-
rangements for Ideal Distillation, Ind. Eng. Chem. Process Des.
Dev., 25, 694699 (1986)
Hernndez, J. G. S., S. Hernndez and A. Jimnez; Analysis of Dynam-
ic Properties of Alternative Sequences to the Petlyuk Column,
Comput. Chem. Eng., 29, 13891399 (2005)
Hernndez, S. and A. Jimnez; Design of Optimal Termally-Coupled
Distillation Systems Using a Dynamic Model, Chem. Eng. Res.
Des., 74, 357362 (1996)
Hernndez, S. and A. Jimnez; Design of Energy-Efcient Petlyuk Sys-
tems, Comput. Chem. Eng., 23, 10051010 (1999)
Jimnez, A., N. Ramrez, A. Castro and S. Hernndez; Design and
Energy Performance of Alternative Schemes to the Petlyuk Distil-
lation System, Chem. Eng. Res. Des., 81, 518524 (2003)
Kidnay, A. J. and W. R. Parrish; Fundamentals of Natural Gas Pro-
cessing, 1st ed., pp. 123, Taylor and Francis, Boca Raton, U.S.A.
(2006)
Knapp, J. P. and M. F. Doherty; Termal Integration of Homogeneous
Azeotropic Distillation Sequences, AIChE J., 36, 969984 (1990)
Krajnc, M. and P. Glavic; Prices of Utilities and Process Structure,
Comput. Chem. Eng., 20, 183188 (1996)
Lee, S. H., M. Shamsuzzoha, M. Han, Y. H. Kim and M. Y. Lee; Study
of the Structural Characteristics of a Divided Wall Column Using
the Sloppy Distillation Arrangement, Korean J. Chem. Eng., 28,
348356 (2011)
Linnhof, B., H. Dunford and R. Smith; Heat Integration of Distillation
Columns into Overall Processes, Chem. Eng. Sci., 38, 11751188
(1983)
Long, N. V. D. and M. Y. Lee; Improved Energy Efciency in Debottle-
necking Using a Fully Termally Coupled Distillation Column,
Asia-Pac. J. Chem. Eng., 6, 338348 (2011)
Long, N. V. D. and M. Y. Lee; Dividing Wall Column Structure Design
Using Response Surface Methodology, Comput. Chem. Eng., 37,
119124 (2012)
Long, N. V. D., S. H. Lee and M. Y. Lee; Design and Optimization of a
Dividing Wall Column for Debottlenecking of the Acetic Acid Pu-
rifcation Process, Chem. Eng. Process., 49, 825835 (2010)
Manley, D. B.; Deethanizer/Depropanizer Sequences with Termal and
Termo-Mechanical Coupling and Component Distribution, U.S.
Patent, No. 5,673,571 (1997)
Mak, J.; Confgurations and Methods for Improved NGL Recovery,
U.S. Patent, No. 7,051,552 (2006)
Malinen, I. and J. Tanskanen; Termally Coupled Side-Column Con-
fgurations Enabling Distillation Boundary Crossing. 1. An Over-
view and a Solving Procedure, Ind. Eng. Chem. Res., 48, 6387
6404 (2009)
Midori, S., S. N. Zheng and I. Yamada; Analysis of Divided-Wall Col-
umn for Extractive Distillation, Kagaku Kogaku Ronbunshu, 26,
627632 (2000)
Montgomery, D. C.; Design and Analysis of Experiments, 3rd ed., pp.
197386, John Wiley & Sons, New York, U.S.A. (1991)
Nejad, S. J., H. Abolghasemi, M. A. Moosavian and M. G. Maragheh;
Fractional Factorial Design for Te Optimization of Supercriti-
cal Carbon Dioxide Extraction of La
3+
, Ce
3+
and Sm
3+
Ions from
a Solid Matrix Using bis(2,4,4-trimethylpentyl)Dithiophosphin-
ic Acid+Tributylphosphate, Chem. Eng. Res. Des., 89, 827835
(2011)
Oluji, ., M. Jdecke, A. Shilkin, G. Schuch and B. Kaibel; Equip-
ment Improvement Trends in Distillation, Chem. Eng. Process.,
48, 10891104 (2009)
Persson, K. and O. strm; Fractional Factorial Design Optimization
of the Separation of Pilocarpine and Its Degradation Products by
Capillary Electrophoresis., J. Chromatogr. B, 697, 207215 (1997)
Peters, M. S. and K. D. Timmerhous; Plant Design and Economics for
Chemical Engineers, 4th ed., pp. 523536, McGraw-Hill, New
York, U.S.A. (1991)
Premkumar, R. and G. P. Rangaiah; Retroftting Conventional Column
Systems to Dividing Wall Columns, Chem. Eng. Res. Des., 87,
4760 (2009)
Schultz, M. A., D. G. Stewart, J. M. Harris, S. P. Rosenblum, M. S.
Shakur and D. E. OBrien; Reduce Costs with Dividing Wall Col-
umns, CEP, 6471 (2002)
Seidel, M.; Verfahren zur Gleichzeitigen Zerlegung von Verfssig-
ten Gasgemischen und anderen Flssigkeitsgemischen mit mehr
als zwei Bestandteilen durch Rektifkation, German Patent No.
294 Journal of Chemical Engineering of Japan
610,503 (1935)
Sinnott, S. K.; Chemical Engineering Design, 4th ed., Coulson & Rich-
ardsons Chemical Engineering Series, Vol. 6, pp. 566568, Else-
vier Butterworth Heinemann, Oxford (2005)
Smith, R.; Chemical Process Design, pp. 405426, McGraw Hill, New
York, U.S.A. (1995)
Tedder, D. W. and D. F. Rudd; Parametric Studies in Industrial Distil-
lation: Part I. Design Comparisons, AIChE J., 24, 303315 (1978)
Triantafyllou, C. and R. Smith; Te Design and Optimization of Fully
Termally Coupled Distillation Columns, Chem. Eng. Res. Des.,
70, 118132 (1992)
Watkins, R. N.; Petroleum Refnery Distillation, 2nd ed., pp. 2760,
Gulf Publishing, Houston, Texas, U.S.A. (1979)

S-ar putea să vă placă și