You are on page 1of 17

GAS TREATING WITH PROMOTED AMINES

David J. Vickery, Susann W. Campbell and Ralph H. Weiland


Taylor, Weiland and Associates, Inc
Technology Development Center
P.O. Box 188, Potsdam, NY 13676

The most important single difference among prim&ry, secondary and tertiary
amines is t h a t tertiary amines do not directly react with carbon dioxide to
form carbamates; at most, they catalyze the hydrolysis of carbon dioxide to
bicarbonate. This simple fact lends several advantages to tertiary amine
treating that have been exploited commercially, particularly with methyldieth-
anolamine (MDEA). The relative chemical kinetics of the hydrogen sulfide and
carbon dioxide reactions make MDEA highly selective for hydrogen sulfide at
short contact times. Like any other alkaline substance, however, MDEA is"
thermodynamically selective for carbon dioxide; thus as contact times
l e n g t h e n , selectivity for hydrogen sulfide gradually weakens. Nevertheless,
good selectivity is one of the most important attributes of MDEA. Another is
the relatively weak binding of carbon dioxide in MDEA solution with the result
that MDEA is more economically stripped than primary and secondary amines.
Use of MDEA for selective removal used to be limited largely to licensed pro-
cesses~ however, generic MDEA is now being used very successfully and with in-
creasing frequency.

ADVANTAGES OF MIXED AMINES

To a degree, the competition between chemical kinet-


Selectivity ics and thermodynamics allows one to exercise some
control over the selectivity of MDEA towards hydrogen
sulfide over carbon dioxide by adjusting the gas-liquid contact time. This
can be done through appropriate choice of tower hydraulics including liquid
depths on trays, solvent circulation rates, packing versus trays and so on.
In fact, the very short liquid residence times in packed towers recommends the
use of packed equipment for increased selectivity when these packed columns
are designed using reliable methods. Higee promises to give degrees of
selectivity well beyond what can be reached using conventional technology.
However, in many applications one finds that MDEA is too selective, i.e., it
allows too much carbon dioxide to slip through the absorber, while diethanol-
amine (DEA) and monoethanolamine (MEA) are not selective enough. Lack of
sufficient selectivity causes too much carbon dioxide to be removed and pushes
up solvent regeneration costs. The best solvent is one with exactly the right
selectivity. This can be produced by blending a primary or secondary amine
with a tertiary amine such as MDEA in just the right proportions.

Paper presented at the 38th Laurance Reid Gas Conditioning Conference, Norman,
Oklahoma, March 7-9, 1988.
Primary and secondary amines absorb carbon dioxide
Promoted Carbon rapidly and produce'very sweet gases; however, their
Dioxide Removal regeneration is very energy intensive. On the other
hand, tertiary amines are relatively inexpensive to
regenerate; the penalty is that they ~bsorb carbon dioxide only very slowly
and the removal is not deep. The advantages of fast absorption, easy re-
generation and deeper removal can be enjoyed simultaneously by using a tertia-
ry amine such as MDEA promoted with relatively small amounts of a primary or
secondary amine. A small amount of DEA is a typical promoter for carbon
dioxide removal by MDEA.

There are other reasons why mixtures of amines might


Solvent Changeouts be encountered in gas processing. For example, car-
ryover of solvent from one process plant to another
and inadvertent solution makeup using the wrong amine will put the solvent in-
to the 'mixed amine' category. Solvent changeouts from one amine to another
are practiced frequently and are generally done by shutting down the gas
plant, draining the old solvent, flushing the system, charging with the new
solvent and startupo Apart from lost production and wasted amine, spent
solutions and plant flushings are classed as hazardous wastes and their dis-
posal has a high associated cost. Unless equipment modifications are necessa-
ry, a solvent changeout can usually be done 'on-the-fly' by making up losses
of the original solvent using the new amine until the plant is operating on
the new solvent alone. In the interim, of course, the plant must successfully
operate over a very wide spectrum of amine mixtures indeed.

Mixed amines represent a range of new gas treating


Blend Selection solvents that can be created by mixing inexpensive,
generic chemicals in the right proportions. One of
the key factors is determining just what the right proportions are for a spec-
ific application. The pilot plant has traditionally been the way that new
s o l v e n t s h a v e moved from the laboratory into the production plant. This is a
very expensive and time consuming route that very few commercial enterprises
can afford. For amine mixtures, the pilot plant is not a feasible route in
the long term because each time another plant is to be fitted w i t h a new
solvent, fresh pilot scale data must be taken. The application of mixed
amines (blends or promoted amines) to the widest possible range of treating
problems depends on the ready and affordable availability of basic engineering
technology to all who want it. Reliance on rules-of-thumb and operating
history is no longer adequate; instead, the technology must be founded on the
physics and chemistry of the process and on sound engineering principles.
GASPLANT-PLUS embraces such a technology.

GASPLANT-PLUS is a fully-flexible flowsheet simulator specifically for hy-


drogen sulfide and carbon dioxide removal by mixtures containing any two of
MEA, DEA, DGA, DIPA and MDEA in water, as well as by mixtures containin 8 a
single amine in water. The unique features of GASPLANT-PLUS are its flowsheet
flexibility, a fundamentally-sound mass transfer rate model for frayed and
packed columns and a thermodynamic package of sufficient rigqr to allow mixed
amines to be handled reliably. We will begin by describing the principles on
which the column model contained within the broader flowsheeting framework of
GASPLANT-PLUS is based. This will be followed by (i) a comparison between
GASPLANT-PLUS predictions and some commercial data on a DEA-MDEA mixture, (ii)
a comparison between a packed and a trayed column for a selective removal ap-
plication and (iii) an example of a solvent changeout without plant shutdown.

AMCOLR - A MASS TRANSFER RATE MODEL FOR AMINE TOWERS

Traditionally, all tower calculations produce either the number of equilibrium


stages needed to perform a specified separation or the separation achieved by
a specified number of ideal trays. The translation from real trays to
equilibrium stages and back again depends on tray efficiency information. It
is enlightening to consider for a moment the situation that would exist if
heat exchangers were designed by this method.
W~

The heatltransfer equivalent of an equilibrium-stage,


Heat Transfer Analogy mass-transfer model produces water at 220F from the
tubes of an exchanger of any type'as long as it has
220F condensing steam on its shell side! To find the real temperature of the
hot water leaving the heat exchanger tubes, a heat exchanger designer using
the equivalent of standard mass transfer design methods would now apply an
efficiency. The efficiency would be characteristic of both the construction
and operating conditions of the particular exchanger. It would depend on the
flow rates and physical properties of the shell-side and tube-side fluids, the
tube diameter, tube length, number of tubes, number of tube-side passes, tube
metal and tube wall thickness, shell diameter, baffle arrangement, number of
baffles and number of shell-side passes as well as the condition of the ex-
changer with respect to scale and dirt. This 'state-of-the-art' mass transfer
approach seems ludicrous when applied even conceptually to heat transfer, yet
it is applied every day in separation process modeling, design and analysis as
a matter of coursel!

AMCOLR takes an approach to design and analysis of


The Rate Approach amine contactors and regenerators that is analogous
to the actual procedures used in heat exchanger de-
sign. In this approach, mass-transfer film coefficients (analogous to the
tube-side and shell-side coefficients of heat transfer) are used alon E with
physical property data and concentration (temperature) driving forces to de-
termine the rates at which individual chemical components (heat) transfer from
phase to phase. The rate approach to modeling any separation process rests on
three basic factors: (i) material and energy balances, (ii) mass and energy
transfer rate models and (iii) vapor-liquid equilibrium models. If the
separation process involves chemical reactions, then there is a fourth factor,
namely, (iv) models for reaction effects on mass transfer. A rate model for
amine regeneration was first developed by us (Weiland and Rawal, 1980~
Weiland, Rawal and Rice, 1981, 1982), then we developed an amine column model
in a more general way (Sardar and Weiland, 1984) and applied it to commercial
operations (Weiland, Sardar and Subramanian, 19857 Sardar and Weiland, 1985).
T h i s model has since been used by Dow Chemical as the technical basis for a
simulator used by their GAS/SPEC Technology Group. A detaile~ description has
been given by Sardar and Weiland (1984) in the context of amine treating and
by Taylor and Weiland (1985, 1987) and Taylor, Vickery and Weiland (1986) in
the context of multicomponent, nonreactive separations in general.
Equilibrium stake methods apply component and energy
Material Balances balance calculations around each complete ideal stake
in a tower. In contrast, AMCOLR does component and
enerKy balances around each phase present on every real stage of the column
(or, in the case of packed towers, in each horizontal slice of the tower).
The individual phase balances are linked by the requirements that (i) the mass
transfer rate of each chemical component, as well as energy, from one phase
must equal the respective transfer rate into the other and (ii) the phases are
in mechanical, thermal and chemical equilibrium at the interface between them.
There is no requirement that the bulk phases themselves be in equilibrium; in
fact, departures from equilibrium are required for mass transfer (i.e., for a
separation) to take place.

Mass and enerEy are transferred from one phase to


Mass and EnerEy another at rates that depend on individual film coef-
Transfer Rates ficients for mass and heat transfer and on concentra-
tion and temperature differences that serve as driv-
ing forces for the transfer° Both the gas and liquid phases are important in
determining the overall transfer rates because both phases offer resistance to
transfer. This approach is completely analogous to the one used in heat
transfer; there, one uses correlations for shell-side and tube-side heat
transfer coefficients alonK with dirt factors and logarithmic mean temperature
differences to calculate heat transfer rates. For mass transfer, there are a
number of reliable mass-transfer-coefficient correlations available for
several typgs of trays, for dumped packing of various sizes, materials and
styles, and for certain structured packinKs° These coefficients depend on
physical and transport properties, flow rates, temperature, pressure and type
of equipment, including such details as the packing type, size, material and
number of redistributors and the tray type and its construction details. The
rate model has been applied very successfully to both trayed and packed
equipment. Its Krea t versatility and general applicability have been de-
monstrated further by our development of a proprietary prototype model for gas
treating w i t h Higee for Fluor Technology, Inc.

Phase equilibrium plays almost as important a role in


Phase E q u i l i b r i u m rate models as it does in equilibrium staEe calcula-
tions. In a rate model, phase equilibrium is assumed
at the fluid-fluid interface, not between the bulk phases leavinE a tray or
packed section. Thus, phase equilibrium is one of the key factors in setting
drivin E forces for mass transfer and, therefore, in determininE the separation
that is actually achieved on a real tray or within an actual section of pack-
inE. The liquid-phase activity coefficient model we use is based on the ex-
tended Debye-Huckel model of Edwards, Newman and Prausnitz (1975) as described
by Deshmukh and Mather (1981)o Because the model is thermodynamically sound,
we have been able to extend it to mixed amine systems° These extensions have
b e e n described in some detail by Chakravarty~ Phukan and Weiland (1985) and
Vickery and Weiland (1985, 1987> and predictions of the thermodynamic model
have been checked aEainst limited experimental data.

4
Liquid-phase chemical reactions always increase mass
Reaction Effects & transfer rates. Reaction effects are usually lumped
Enhancement Factors into so-called enhancement factors. Unlike tray ef-
ficiencies and HETPs, enhancement factors are not em-
pirical constants to correct for the effects of chemical reaction. They are
very well defined parameters that are calculated from a knowledge of the
chemical kinetics, the reaction stoichiometry, liquid composition and a few
physical and transport properties of the components in the liquid.
Enhancement factors depend, therefore, on such parameters as temperature,
amine concentrations, acid gas loadings, kinetic constants, and so on. We
have developed reaction models that correctly account for the way that two
amines interact with each other during the process of absorbing the acid
gases. These models are not empSrical; they are based on knowledge of the
actual kinetics involved.

The rate model for amine columns, AMCOLR, is central


Flowsheeting and to the general utility and high reliability of the
Other Features GASPLANT-PLUS flowsheet simulator. GASPLANT-PLUS has
been designed to allow complete flexibility in creat-
ing gas processing flowsheets. For example, single flow and split flow trains .
of several absorbers and regenerators can be handled along with associated
heat exchange equipment and flash tanks with or without flash gas treaters.
Towers can be specified by actual diameter or percent of flood, they can be
swaged up or down wherever and as often as desired and tray types or packing
types can vary from place-to-place within the towers. The absorption of hy-
drocarbons and other gases is calculated and amine losses are determined.
Calculations for amine mixtures can be done on mainframe and m i n i c o m p u t e r s us-
ing G A S P L A N T - g L U S and on personal computers using GASPLANT-PLUS/PC.

COMMERCIAL DATA - CODy~ W Y O M I N G PLANT OF MARATHON OIL COMPANY

At the 1987 Laurance Reid Gas Conditioning Conference some performance data
from Marathon's Cody, Wyoming plant were presented by Harbison and Handwerk
(1987). This plant, built in 1985, normally employs a generic MDEA system at
a design c o n c e n t r a t i o n of between 25~ and 30~. However, Harbison and Handwerk
reported that solution analysis obtained during plant tests conducted during
mid-1986 showed a solvent containing about 22% MDEA and 4Z DEA. The presence
of the DEA was unintent.ional and may have resulted from DEA carryover into the
MDEA circuit, or inadvertently making up DEA to the MDEA system. As reported
in their paper, the presence of DEA was a nuisance. For our purposes, howev-
er, the presence of DEA was a great stroke of luck! This plant was actually
operating with a DEA-MDEA blend at the time of the tests and, fortunately, the
test data were given in sufficient detail to satisfy the data requirements of
GASPLANT-PLUS.

The tests appear to have been run on the amine contactor and the regenerator
separately because the test data for these two units give different solvent
strengths (21.8 wtZ MDEA and 4.2 wt~ DEA for the contactor versus 18.2~ MDEA
and 4.3% DEA for the stripper). There are several inconsistencies in the
plant data but none are serious enough to prevent comparisons with simulated
performance; nonetheless, they are worth pointing out. Gas analyses indicated
that the net carbon dioxide and hydrogen sulfide pickup rates in the absorber
were 38.9 and 13.4 ibmol/hr, respectively; liquid analyses gave correspondin 8

5
figures of 46.0 and 16.0 ibmol/hr. Thus material balances around the con-
factor were off by between 18% and 20%. The acid gas loads in the lean
solvent from the stripper are reported to be exactly 0.02 for carbon dioxide
and 0.004 for hydrogen sulfide and the stripper material balance data given in
their Table 5 give 100% closure on all components; there is no indication as
to what was measured and what was calculated.

The amine contactor data used in the GASPLANT-PLUS


Amine Contactor simulation included the tower ID (3 ft), tray type
(V-I valve), actual number of trays (i0), active tray
area (5.135 sq ft), weir height (2 in) and tray spacing (2 ft). The rest of
the data provided in Table 4 of the paper were either already contained as
part of tray correlations within the GASPLANT-PLUS data base or they were
calculated as part of the simulation itself (e.g., flows, residence times,
etc.). It was assumed that the stated lean preloads of 0.02 for carbon di-
oxide and 0.004 for hydrogen sulfide and the sour gas flow and analysis were
correct. Measured and simulated contact0r performance data are shown in Table
1 along with performance predictions for straight MDEA.

TABLE 1 Contactor Performance

Plant DEA-MDEA 25%


Data Mixture MDEA

Sweet Gas:

Mole% C02 14.38 13.60 14.07


ppm H2S 220 558 233

Rich Solvent:

C02 Load 0.46A 0.5030 0.&936


H2S Load 0.157 0.1554 0.1675
Temperature (F) 139 142.1 141.1

Temperature Bulge:

Tray Number 2 3 3
Liquid (F) 139 153.4 150.4

The MDEA system was designed to provide a satisfactory feed to a sulfur recov-
ery~ unit, with the product 'sweet' gas being further treated in a DEA unit.
Thus, the goal was bulk hydrogen sulfide removal, not production of 1/4- grain
gas. Solvent could be fed to any of trays 6, 8, 10, 11 and 14 of the 14-tray
absorber; the results reported in Table 1 above are for feed to the 10th tray.
Calculated, tray-to-tray profiles of moleX hydrogen sulfide in the gas are
shown in Figure i as a f u n c t i o n of the number of trays being used in the con-
tactor. It is apparent from the plot that with 10 active trays in the absorb-
er, the bottom part of the column limits the degree of hydrogen sulfide re-
moval. The bottom four trays are doing worse than no hydrogen sulfide pickup
at all - two of them are doing some stripping as a result of continued carbon
dioxide removal. Column p e r f o r m a n c e under such • circumstances is very
sensitive to operating conditions; for example, decreasing the lean
temperature by 5F can lower the hydrogen sulfide content of the sweet gas by a
factor of nearly two.

For lack of information on the physical dimensions of


Stripper the stripper and the trays it contained, we have as-
sumed that • it is of identical construction to the ab-
sorber. The stripper contained 22 trays and rich solvent entered on the 18th
tray (from the bottom). From flowsheet data, the reboiler heat load was calc-
ulated to be 4.647 MMBtu/hr) however, a 5 . 0 0 1 M M B t u / h r reboiler heat load was
found necessary to match the stated reflux flow of 7&.~5 ibmol/hr. The summa-
ry results of Table 2 show reported operating data compared with GASPLANT-PLUS
simulation data for both reboiler heat loads. The offgas composition shows
very little sensitivity to heat input to the reboiler because the solvent in
both cases is stripped so cleanly. The hydrogen sulfide and carbon dioxide
contents of the lean, though, do depend on reboiler duty. Predicted lean
loads are well below reported values; however, the reported data may represent
only limits of measurability.

PACKING VERSUS TRAYS FOR SELECTIVE REMOVAL

The selectivity of MDEA or any other amine for hydrogen sulfide is higher the
shorter the contact time between the phases. It is this factor that makes gas
treating using Higee (Bucklin and Won, 1987; Buckingham and Bucklin, 1988) so
promising. Compared with trays, however, packed columns have rather small
liquid holdup volumes. Lower holdup gives packed equipment an advantage over
trays in certain selective removal applications; however, this advantage is
not being exploited today. One of the reasons may be lack of sufficient ex-
perience with packings in gas treating applications; another is probably the
lack of reliable design methods for packed equipment in general. The GASPLANT
family of amine treating software allows packed columns to be designed and
simulated with the same high degree of reliability as trayed equipment.

P~
Field data on a 14-tray absorber processing natural
Performance of a gas at 915 psia form the base case for this example.
Packed Absorber The .solvent was 50% MDEA at 240 gpm and the sour gas
contained 4.07 and 0.4% carbon dioxide and hydrogen
sulfide, respectively. Field measurements showed a sweet gas having between 8
and 12 vppm hydrogen sulfide and 1.8% to 2.2% carbon dioxide, i.e., between
• . . . - ............. . . . . , :. . . . . . . . . . . . . . . . . . . . . . ~ . . . . . . . . . . . . •

TABLE 2 Stripper Performance - Component Flows (ibmole/hr)

Test Reboiler Duty Reboiler Duty


Data 4.65-MMBtu/hr 5.00 M M B t u / h r

Off Gas:

H20 4. 646 4.803 4.809


DEA - 0o 0 0.0
MDEA - 0.0 0.0
H2S 13.41 13. 704 13. 701
C02 38.43 39. 231 39. 306
CH4 0.15 0. 149 0. 149
C2H6 0.025 0.025 0.025
C3H8 0. 016 0. 016 0. 016
N2 0.046 0.046 0.046
Total 56.7 57. 739 57. 817

Lean Solution:

H20 1,617.2 1,617.2 1,617.2


DEA 15.7 15.7 15.7
MDEA 57.5 57.5 57.5
H2S 0.29 0.00043 0.00025
C02 1.46 0.678 0.599
CH4 - 0.0 0.0
C2H6 - 0.0 0°0
C3H8 - 0.0 0.0
N2 - 0.0 0°0
Total 1,692.1 1,691.08 1,691.00

45% and 55% carbon dioxide slip. A l t h o u g h blend capabilities were not needed
for this straight MDEA case, GASPLANT-PLUS was used for the simulation.
Actual performance v s . - s i m u l a t i o n data for the trayed absorber are shown in
Table 3 from w h i c h it can be seen that GASPLANT-PLUS predicts the performance
of this absorber very accurately. Table 3 also shows the predicted per-
formance of a column containing 30 feet of 2 inch steel Pall rings. The col-
umn was sized for 75% of flood but was otherwise simulated w i t h operating con-
ditions identical to the actual 14-tray absorber, i.e., the same solvent flow,
preload, etc. The packed column slips 75% of the carbon dioxide (vs. 45% for
trays) and this greatly reduced carbon dioxide pickup allows better than
i/4-grain gas to be produced. The lower carbon dioxide pickup will also allow
sufficiently stripped solvent to be produced using less heat to the re-
generator. It is apparent from this example that a packed absorber maximizes
carbon dioxide slip, it results in lower energy requirements and .it allows
better hydrogen sulTide removal.
i

. . . . . . , . : . . . . . . , . . ,

TABLE 3 Trays vs. Packing in Selective Removal Application

Field 14 Valve 30-ft


Data Trays Pall Rings

Sweet Gas:

H2S (ppm) 8 - 12 11.2 2.57


C02 (mole%) 1.8 - 2.2 1.85 3.05

C02 Slip (%) 45 - 55 45.1 75.2

Temperature Bulge: mid-tower mid-tower


171F 146 F

Rich Solvent:

C02 Load 0.597 0.273


H2S Load 0.1094 0.1094
L

Total 0.706 0.482

SOLVENT CHANGEOUTS WITHOUT PLANT SHUTDOWN

For reasons of greater selectivity, lower regeneration energy requirements


through reduced circulation rates, less corrosivity and possibly greater amine
stability, MDEA continues to replace DEA in numerous selective hydrogen
sulfide removal applications. Solvent changeouts from DEA to MDEA often re-
quire no equipment modifications, yet they are generally achieved by shutting
down the plant, draining the old solvent, cleaning, and finally recharging
with MDEA.

Gradual solvent changeouts have the advantages of no lost production, no dis-


posal problems with the environmentally-hazardous old solvent, no use and sub-
sequent disposal of cleaning agents, and no additional manpower requirements.
An advanced flowsheet ~imulation capability can be used to determine if such a
procedure is feasible and, when it is, plant simulation can help to ensure
that the solvent changeout is done reliably and with no production or cost
penalties.

The data in this example are taken from an operating


Original DEA Plant DEA plant for which a conversion to 30% MDEA is plan-
ned. The plant processes 50 MMscfd of sour gas con-
taining 12% hydrogen sulfide and 5% carbon dioxide at I000 psia and makes
i/4-grain gas using a solvent flow of 900 gpm of 30% DEA to the absorber.
The absorber and regenerator contain 22 siev~ trays each, with reflux to the
third tray (from the top) of the stripper. The reboiler duty is 45 MMBtu/hr
and the rich solvent from the absorber is flashed at 50 psia. Gradual con-
version to 30% MDEA is desired. Some GASPLANT-PLUS predictions of plant per-

9
TABLE 4 Simulated Plant Performance with DEA and MDEA

30 w t% DEA 30 wt% M D E A

Sweet Gas:

C02 (vppm) 0.68 2800.


H2S (vppm) 3.01 0.28
J

Rich Solution:

C02 (mol/mol) 6.2240 0.2340


H2S (mol/mol) 0.5162 0.5785

Lean Solution:

C02 (mol/mol) 0.0143 0°0045


H2S (mol/mol) 0.0132 0.~009

formance are shown in Table 4 for operation using 30% DEA along with corres-
ponding figures for 30% MDEA. Both sets of data are based on identical
operating conditions; all that has been changed is the solvent. GASPLANT-PLUS
accurately matches the 4 vppm specification on hydrogen sulfide and, as is
quite typical of DEA plants, the sweet gas contains very little carbon
dioxide.

GASPLANT-PLUS predicts that when the plant is switch-


Plant Operation ed to 30% MDEA the same reboiler load of A5 MMBtu/hr
with 30 wt% MDEA will produce gas much lower in hydrogen sulfide and,
at the same time, significant carbon dioxide can
be slipped. (However, rich solution loads with respect to carbon dioxide are
higher because of the lower molar concentration of MDEA.) The MDEA is somewhat
overstripped with a &5 MMBtu/hr heat flow to the reboiler and, in fact, 1/4
grain gas containing about 3700 ppm carbon dioxide could be produced with
about 10% lower circuletion and 10% lower reboiler heat load.

Now that we have established satisfactory performance


Gradual ChanEeouts: for the DEA plant operated using 30% MDEA, it remains
Plant Operations with only to show that a smooth, Eradual transition can be
Mixtures of Amines made from one to the other. GASPLANT-PLUS predictions
of sweet gas purities and the lean solvent loadinEs
are shown in FiEures 2 and 3 as a function of blend composition for a constant
total amiue concentration of 30~° It is immediately obvious that as far as
producing I/A grain Eas is concerned the chanEeout from DEA to MDEA can indeed
be done g r a d u a l l y . Unless other, overriding factors are involved, there is
absolutely no need to Eo through the expense of a plant shutdown. The simula-
tions shown in Figures 2 and 3 also warn of when the Ereatest changes in per-
formance are likely to be felt. For example, they indicate that as the sol-
vent approaches s t r a i g h t MDEA, some significant increases in both carbon

10
dioxide slip and hydroEen sulfide pickup (i.e., selectivity) will be seen.
These altered sweet gas compositions also indicate that it is towards the low
DEA end of the mixture that reboiler heat load can begin to be reduced
significantly. There are several striking observations to be noted:

(i) all other operating conditions remaining the same, the use of a 30%
DEA-MDEA blend containing only 5% DEA lowers the sweet Eas carbon di-
oxide content by more than a factor of ten below 30% MDEA alone.

(ii) it is easier to strip a 5% DEA-25% MDEA mixture than any other combina-
tion, including the pure components,

(iii) the use of small amounts of DEA as a promoter is a very effective way
to tailor carbon dioxide pickup by MDEA without greatly affecting hy-
droKen sulfide removal.

The ease w i t h w h i c h carbon dioxide can be stripped from 30% DEA-MDEA blends in
this particular plant is, in fact, due to the operation of the stripper
itself. FiKure 4 is a plot of tray-to-tray loading profiles of carbon dioxide
in the regenerator with percent DEA in the total 30% blend as parameter. Rich
solvent, which enters on tray 3 (from the top), always has m u c h the same .
carbon dioxide loading regardless of the amine mixture involved. However, as
small amounts (0.5%, I%, etc.) of DEA are added to the mixture w i t h MDEA, the
extent of stripping for the same reboiler heat input increases dramatically
until the solution contains about 5% DEA. Further additions of DEA then cause
a reversal and stripping becomes poorer; we find that eventually straight DEA
is somewhat harder to strip than MDEA of the same w e i g h t percent.

The dramatic, positive effect of small amounts of DEA and the ultimate re-
versal of this effect as DEA concentrations continue to be increased is the
result of competition between (i) the kinetics of the process by w h i c h carbon
dioxide is transported out of the liquid into the vapor, and (ii) the
thermodynamics of equilibrium between the phases. The chemical reaction
between carbon dioxide and DEA to form a carbamate allows the carbon dioxide
to be piggybacked to the interface where it is released as free dissolved
gas.The carbon dioxide - DEA reaction tends always to enhance the mass transf-
er process; the larger the amount of DEA, the greater the enhancement. As we
have shown elsewhere (Chakravarty, Phukan and Weiland, 1985; V i c k e r y and
Weiland, 1986, 1987), the presence of DEA lowers equilibrium pressures of
carbon dioxide above loaded solutions compared to the backpressures over cor-
responding solutions of MDEA alone. Again, the more DEA there is, the lower
the backpressures are. These effects compete to produce an optimum blend as
far as ease of stripping is concerned. When very small amounts of DEA are
present in the MDEA, mass transfer kinetics are Kreatly enhanced; when the
amount of DEA is increased beyond the optimum level, however, the negative
effects of reduced equilibrium partial pressures start to dominate. We
believe that this is a general effect and should be observed in most cases in-
volving mixed amine systems.

CONCLUSION

Blended or mixed amine technology and the availability of reliable flowsheet


simulation capabilities opens up some interesting new ways of processing
gases. It becomes possible to tailor the amine mixture so that a sweet gas

II
• .

• . -

containing just the right amounts of both carbon dioxide and hydrogen sulfide
is produced with minimum regeneration energy requirements. In effect, the num-
ber of possible solvents becomes virtually limitless; the 'best' solvent will
be different for each and every gas stream being treated.

The ability to handle mixed amine systems and the high reliability of
GASPLANT-PLUS are a result of the fundamentally sound engineering principles
embodied in its mass transfer rate models and its thermodynamic database.
GASPLANT-PLUS is the only commercial amine plant simulator w i t h a complete
range of mixed amine as well as single amine capabilities, all within a
fully-flexible flowsheeting environment.

The ability to formulate solvent mixtures puts a great de@l of flexibility in-
to the optimization of plant operations. The blend formula or mix is a
parameter of great utility in tuning the selectivity of the solvent.
GASPLANT-PLUS and GASPLANT-PLUS/PC now give every gas processor the ability to
f o r m u l a t e his own solvent using generic materials.

REFERENCES

Buckingham, P. A., Bucklin, R.W., "Test Results of Higee Performance in


Selective Hydrogen Sulfide Removal Service", Proceedings of the 38th Laurance
Reid Gas Conditioning Conference, Paper ' M ' , Norman, Oklahoma, March 7-9,
1988.

Bucklin, R. W., Won, K. W., "Higee Contactors for Selective Hydrogen Sulfide
Removal and Superdehydration", Proceedings of the 37th Laurance Reid Gas
Conditioning Conference, DI-DI6, Norman, Oklahoma, March 2-A, 1987.

Chakravarty, To, 9hukan, U. K., Weiland, R. H., Chemical Engineering Progress,


p.32, April, 1985.

Deshmukh, R. D., Mather, A. W., Chem. Eng. Sci., 36, 355 (1981).

Edwards, T. J., Newman, J., Prausnitz, J. M., AIChE Journal, 21.1, 248-259
(1975).

Harbison, J. L., Han4werk, G. E., "Selective Removal of Hydrogen Sulfide


Utilizing Generic MDEA", Proceedings of the 37th Laurance Reid Gas Condition-
ing Conference, HI-H12, Norman, Oklahoma, March 2-&, 1987."

Sardar, H., Weiland, R. H., "A Non-equilibrium Stage Approach to the Design
and Simulation of Gas Treating Units", AIChE Annual Meeting, San Francisco,
California, November 25-30, 198&.

Sardar, H., Weiland, R. Ho, "Simulation of Absorbers and Strippers in


Commercial Amine Treating U~its", AIChE Spring National Meeting; Houston,
Texas, March 24-28, 1985.

Taylor, R., Vickery, D. J., Weiland, R. H., "A Transfer Rate Based Module for
the Computer-Aided Simulation of Multicomponent Distillation and Absorption
Operations", Canadian Chemical Engineering Conference, Sarnia, Ontario, Octob-
er, 1986.

12
Taylor, R., Weiland, R. H., "The Rate Approach to Separation Process Simula-
tion", European Federation of Chemical Engineers, Working Party on Absorption,
Distillation and Extraction, Amsterdam, June 6, 1985.

Taylor, R., Weiland, R. H., Proceedings of XVIII Congress: The Use of


Computers in Chemical Engineering, Giardini, Naxos (Sicily), Italy, p.&79
(1987).

Vickery, D. J., Weiland, R. H., "Solubility of Acid Gases in Blends of


Amines", AIChE Annual Meeting, Miami Beach, Florida, November 2-7, 1986.

Vickery, D. J., Weiland, R. H., "Simulating Gas Treating with Blended Amines",
AIChE Spring National Meeting, Houston, Texas, March 29 - April 2, 1987.

Weiland, R. H., Rawal, M. Yo, "Stripping of Carbon Dioxide from Rich


Monoethanolamine Solutions", Proceedings of Chemeca 80, Melbourne, Australia,
f~
1980.

Weiland, R. H., Rawal, M. Y., Rice, R. G., "Stripping of Carbon Dioxide from
Monoethanolamine in Continuous Contact Equipment", AIChE Annual Meeting, New
Orleans, Louisiana, November, 1981.

Weiland, R. H., Rawal, M. Y., Rice, R. G., AIChE Journal, 28, 963-973 (1982).

Weiland, R. H., Sardar, H., Sivasubramanian, M. S., "Simulation of Commercvial


Amine TreatinE Units", ProceedinEs of the 35th Laurance Reid Gas ConditioninE
Conference, Jl-J43, Norman, Oklahoma, March 4-6, 1985.

13
14
/
/0.278
/
n"
W 12 0.295
O3 /
/
0
rn 10 0.313
$
Z I
I

O9
8 0.339.L-..
>-
< I
rc
F- 61" 0.3
LL \
O %
£E
CLAUS FEED
LU
4 H2S/CO2 = 0.570"

E3
Z 2

0
0.01 0.1 1.0 10.
Mole % H2S
FIGURE I Profiles of H2S Concentration in the Gas as a Function o{ the Number of Trays in the Absorber
Parameter on Figure is the mole r a t i o H2S to C02 in Claus Plant Feed vs. Number of Trays
II I I

I I I I I

3000 3

H2S
E
CI. CI
o.
> 2000 2 oi
>

©
0 7"
O9
< <
C~
I-- l--
W 1000 IJJ
W W
09 09

0 0
0 5 10 15 20 25 30
w t % DEA

FIGURE 2 Sweet Gas P u r i t y as a Function of Biend Composition during Gradual


Solvent Changeout to 30 wt% MDEA

15
I I I I I

0.015 m

O H
E
!

O 0.010
E

£3
<
©
_J
Z
<
LU
..J
0.005

0
0 5 10 15 20 25 30
wt% DEA

FIGURE 3 Lean Solvent Loadings as a Function of Blend Composition during


Gradual Solvent Changeout

15-A
1.0 mmm I I I I I I
I
m
m
m

mm m

I
m

mm

0.1

(D
u
O
E
(D
m
O wt% DEA
E
30
£3
< 0
© 0.01
_.1
OW

C.)
0

.#'~
1

0.001 ~.
3 6 9 12 15 18 21
TRAY NUMBER (from bottom)

FIGURE 4 Tray-to-Tray Loading Profiles of C02 in the Regenerator as a Function


of Blend Composition during Gradual Solvent Changeou% %o 30 w%% MDEA

16

You might also like