Anda di halaman 1dari 16

Jefftreat® Product Technical Note

abcde
Amine Degradation Chemistry in CO2 Service

presented at the

48th Annual

Laurance Reid Gas Conditioning Conference

March 1-4, 1998

Norman, Oklahoma

By:

Patrick E. Holub, P.E.


Dr. James E. Critchfield
Dr. Wei-Yang Su
Huntsman Corporation
3040 Post Oak Blvd.
Houston, TX 77056

Abstract
This paper describes the degradation chemistry of ethanolamines in CO2 service, the type of
degradation products formed, and a proposed formation mechanism for the degradation products.
Laboratory data and plant sample analysis are presented that show amine degradation and its
effects on the measured concentration of metals in the solution.

Keywords: carbon dioxide, corrosion, degradation products, diethanolamine, ethanolamine


blends, gas treating, methyldiethanolamine, methyl-monoethanolamine,
monoethanolamine, solvent degradation.

48th Annual Laurance Reid Gas Conditioning Conference March 1-4, 1998
C:\GTData\Gas Treating Literature\Jefftreat Tech Notes\ds115698.doc Page 1
INTRODUCTION

Monoethanolamine (MEA) and diethanolamine (DEA) are both commonly applied for removal of
CO2 from gas streams. Numerous studies have been made over the years to determine the
potential of MEA and DEA to degrade in this service. Most of the information and data
previously presented has focused on laboratory studies of degradation chemistry of the amines in
the presence of CO2 and/or H2S. It is necessary to conduct laboratory testing at elevated
temperatures and high partial pressures of acid gas in order to complete experiments in a
reasonable timeframe. However, these conditions are not maintained for long periods in
operating plants, so the applicability of the data has occasionally been questioned.

Kohl & Nielsen point out one major difference between laboratory results and plant experience.
Under accelerated laboratory conditions, DEA degradation is significantly faster than that of
MEA, however in actual plant operations, the reverse appears to be the case. Evidently,
differences in process conditions may be as important as differences in degrees of amine
degradation. In the reboiler (the point of highest temperature) of a plant using DEA, the solution
is normally well stripped and comparatively free of CO2. For MEA systems, the lean solvent
may contain substantial quantities of CO2 due to lower stripping efficiency. Such factors are
significant, because in laboratory experiments, Kim and Sartori (1984) demonstrated that the rate
of degradation of DEA depends on the amount of CO2 present.

Analysis of degradation products in plant solutions is not simple, and the procedure normally
requires sophisticated (and expensive) analytical methods, such as GC/MS, NMR, and/or LC. Field
analysis generally serves only the purpose of maintaining the amine strength, the depth of stripping,
and adequate circulation rate. Measuring only alkalinity and acid gas content can accomplish these
goals. Unfortunately, these simple measurements tell nothing about the state of amine degradation.
It is common to find substantial degradation of an MEA- or DEA-containing solvent without the
plant staff being aware of the problem.

Industry has long accepted the problems associated with CO2 degradation of MEA solutions. As far
back as March 1957, Lang and Mason presented data showing corrosivity of MEA and MEA
degradation products in CO2 service. This coupled with higher component vapor pressure of MEA
greatly reduces the practicality of using MEA as a formulating agent. For these reasons, MEA
degradation has not been the focus of our investigation.

With the advent of formulated MDEA solvents, the issue of solvent degradation has become a
necessary factor in solvent selection. A considerable amount of recent industry research has been
devoted to demonstrating benefits of solvent formulation with ethanolamine/MDEA blends. In
particular, DEA/MDEA blends have received considerable attention in academic work. A
number of companies have attempted to use DEA/MDEA blends, with varying degrees of
success. This paper will demonstrate from process sample results that formulation with
DEA/MDEA blends does not solve the degradation problems experienced in DEA-only
applications. In contrast, this paper will also demonstrate with laboratory and plant data that
Huntsman Corporation’s JEFFTREAT® M-500 Solvents do not experience these same
degradation problems.

This paper will review our understanding of secondary ethanolamine degradation chemistry as it
pertains to operating plant conditions, provide operating plant and solvent history, and offer
suggestions to minimize the potential for operating problems due to solvent degradation.

48th Annual Laurance Reid Gas Conditioning Conference March 1-4, 1998
C:\GTData\Gas Treating Literature\Jefftreat Tech Notes\ds115698.doc Page 2
ETHANOLAMINE DEGRADATION CHEMISTRY

Certainly, degradation of ethanolamines is not a new topic in gas treating. Over the past 40 years,
there has been considerable study of the degradation of the ethanolamines in this service. MEA and
DEA degradation have been studied at stripper temperatures, and in the presence of acid gases. A
number of possible mechanisms have been proposed for the degradation process. It is our belief
that the mechanism for DEA degradation presented by Kim and Sartori (1984) and by Hsu and Kim
(1985) best reconciles the observations we have made in our own research and in our analysis of
plant samples from CO2 applications.

Dawodu and Meisen (1996) studied degradation of DEA/MDEA and MEA/MDEA blends in the
laboratory. They concluded that both MEA and DEA degrade in the blends. They also concluded
that plants, which use such blends, should be equipped to monitor degradation to minimize
operational problems. They suggested that these amines might degrade sufficiently that makeup of
DEA or MEA could be required in field application of the blends. However, we believe that the
accumulation of the degradation products in these types of blends is the more important process
consequence than the possible loss of capacity due to degradation.

Under CO2 plant conditions, secondary ethanolamine degradation is likely to proceed through the
following steps:

1. amine + CO2 ↔ oxazolidone


2. amine + oxazolidone → ethylenediamine
3. ethylenediamine + oxazolidone → “polymeric” ethyleneamines, and
4. ethylenediamine or higher ethyleneamine → piperazines

Steps 3 and 4 are simultaneous, and step 4 can be considered a termination step in the series
reactions. Steps 1 and 2 are in agreement with Kim and Sartori, as is step 4. We believe that the
reaction of ethylenediamine with oxazolidone to form higher ethyleneamines (Step 3) is likely to be
preferred under CO2 plant conditions over condensation reactions (Step 4).

Oxazolidone Formation

When the ethanolamines react with CO2, the main reaction product is an ethanolamine carbamic
acid. The carbamic acid of the primary and secondary ethanolamines, over time and in the presence
of heat, will undergo further reactions that lead to irreversible degradation of the ethanolamines. The
first step in degradation of secondary ethanolamines (DEA and MMEA) is a reversible reaction with
CO2 to form 5-member ring compounds (oxazolidones):

Carbamic Acid
R R O

| | ||
HO − CH 2 − CH 2 − N − H + CO2 HO − CH 2 − CH 2 − N − C − OH
ethanolamine + CO2 = ethanolamine carbamic acid

where
R = CH3 (MMEA)
R = CH2CH2OH (DEA)

Oxazolidone The first step in ethanolamine degradation (reversible)

48th Annual Laurance Reid Gas Conditioning Conference March 1-4, 1998
C:\GTData\Gas Treating Literature\Jefftreat Tech Notes\ds115698.doc Page 3
ethanolamine carbamic acid - water = an oxazolidone

R O CH 2CH 2


| || / \
OH − CH 2 − CH 2 − N − C − OH O N− R + H 2O
\ /
C
||
O
Substituted Ethylenediamine Formation

With sufficient heat and stripping efficiency, oxazolidone compounds may revert to the parent
ethanolamine. Unfortunately, these same conditions also cause oxazolidone compounds to react
irreversibly in the second step of ethanolamine degradation; that is the formation of substituted
ethylenediamine compounds, such as tris-hydroxyethylethylenediamine (THEED), or dimethyl-
hydroxyethylethylenediamine (DMHEED).

The second step in ethanolamine degradation (irreversible)


ethanolamine + an oxazolidone = a substituted ethylenediamine

R CH 2CH 2 R R

| / \ | |
HOCH 2 CH 2 − N + O N− R HOCH 2 CH 2 − N − CH 2 CH 2 − N + CO2
| \ / |
H C H
||
O
When Parent Ethanolamine Substituted Ethylenediamine
R = CH3 MMEA DMHEED
R = CH2CH2OH DEA THEED

The secondary nitrogen atom in the ethylenediamine is reactive. It can undergo further
decomposition reactions, but once the ethylenediamine has formed, the degradation of the parent
amine is irreversible. In addition, steric hinderance can also play an important role in substituted
ethylenediamine formation.

Substituted Piperazine Formation

The next reaction in the degradation process is desirable, in that it can convert the ethylenediamine
compound into a less-objectionable piperazine. For instance, when DEA is degraded into THEED
in the laboratory, a substantial portion of the THEED can be converted to bis- (hydroxyethyl)
piperazine (bis-HEP). The analogous compound for MMEA is dimethylpiperazine (DMPIP).

substituted ethylenediamine carbamic acid - H2O = a substituted piperazine


R CH 2CH 2
| / \
HO − CH 2 − CH 2 − N − CH 2 − CH 2 − N − R
| → R- N
\
N − R + H 2O
/
H Condensation
CH CH 2 2
When Parent Ethanolamine Substituted Ethylenediamine Resulting Piperazine
R = CH3 MMEA DMHEED DMPIP
R = CH2CH2OH DEA THEED bis-HEP

48th Annual Laurance Reid Gas Conditioning Conference March 1-4, 1998
C:\GTData\Gas Treating Literature\Jefftreat Tech Notes\ds115698.doc Page 4
When the parent ethanolamine is secondary, the resulting piperazine contains no primary or
secondary nitrogen atoms, so it does not react quickly with CO2. However it does act as a base and
it will provide some CO2 carrying capacity. Bis-HEP is thought to be benign with respect to
corrosion in gas treating systems. Blanc reported that the piperazine degradation product in solution
in fact had lower corrosion rates than a solution without the piperazine degradation product.
Unfortunately, Blanc’s study was conducted with H2S and not CO2 and the degraded solution
apparently did not contain any of the precursor degradation compounds, only bis-HEP.
Additionally, the bis-HEP does not undergo further degradation reactions with CO2 nor is it
reversible to the parent degradation molecules. In that sense, bis-HEP is a desirable endpoint in
DEA degradation chemistry.

Unfortunately, analysis of numerous field applications shows that in practice only a minor portion
of THEED converts into bis-HEP. For example, consider the following analysis from a
DEA/MDEA formulation that was in use for CO2 removal in a hydrogen production application.

TABLE 1
Analysis of a DEA/MDEA Formulation For THEED & bis-HEP
MDEA, % Weight 19.1
DEA, % Weight 14.6
THEED, % Weight 9.1
Bis-HEP, % Weight 0.6

Higher Molecular Weight Ethyleneamines

The secondary nitrogen atom in substituted ethylenediamines can also react with oxazolidone; this
allows the ethylenediamine compounds to “chain out” (i.e., to form higher molecular weight
ethyleneamines). It is possible that other routes exist to form polymeric ethyleneamines, particularly
when catalysts are present. One such likely route is the condensation reaction. The secondary
nitrogen in the diethylenetriamine should have similar reactivity towards CO2 as that in the
ethylenediamine. Based on their structure alone, one would suspect that these long-chain amines
would play a role in foaming, fouling and other operating problems.

R R CH 2CH 2 R R R

| | / \ | | |
HOCH 2 CH 2 − N − CH 2 CH 2 − N + O N− R HOCH 2 CH 2 − N − CH 2 CH 2 − N − CH 2 CH 2 − N + CO2
| \ / |
H C H
||
O

a substituted ethylenediamine + an oxazolidone = a substituted diethylenetriamine

These long-chain compounds are difficult to measure since they are unlikely to be recovered
effectively in gas chromatography. Hsu and Kim (1985) reported the presence of triamine
compounds in degraded ethanolamine solutions. Analysis by NMR of the degraded DEA/MDEA
solution presented in Table I indicate that these long-chain compounds are indeed present, however
they are present in smaller concentration than THEED itself.

48th Annual Laurance Reid Gas Conditioning Conference March 1-4, 1998
C:\GTData\Gas Treating Literature\Jefftreat Tech Notes\ds115698.doc Page 5
Laboratory Data

FIGURE 1 shows the results of Huntsman’s laboratory testing of product stability of three different
amine solvents in the presence of CO2; a DEA/MDEA formulation, a MMEA/MDEA formulation,
and Huntsman Corporation’s JEFFTREAT® M-510 solvent. The MMEA and the DEA in the
ethanolamine containing formulations are partially converted into substituted ethylenediamine,
piperazine, and oxazolidone compounds during the tests. The loss rate of MMEA was similar to
that of DEA in these tests. In contrast, JEFFTREAT® M-510 solvent was comparatively stable
under these conditions.

JEFFTREAT® M-500 Series solvents are not formulated with primary or secondary ethanolamines.
Consequently, JEFFTREAT® M-500 solvents are not subject to the same type of degradation
chemistry that occurs in CO2 removal applications utilizing ethanolamines. This results in superior
product stability.

FIGURE 1. Comparison of Stability of JEFFTREAT® M -510


Promoter vs. 2' Ethanolam ines in Formulated Solvents
Accelerated Degradation Tests in a Laboratory Autoclave
120
Promoter in JEFFTREAT® M -510 Solvent
100
Remaining in Solution
% of Active Agent

80

60

M M E A in M M E A /MDEA Formulation
40

20
DEA in DEA/MDEA Formulation

0
0 1 2 3 4 5 6 7
Elapsed Time, Weeks
Under laboratory conditions, no amine solvent appears to be completely immune to degradation.
Although Huntsman’s research has demonstrated that the promoter in JEFFTREAT® M-510
solvent is far more stable than the ethanolamines, under severe conditions the promoter can slowly
react to form an inert compound. At stripper temperature and in the presence of substantial
concentration of CO2, the rate of formation of this inert compound is about an order of magnitude
slower than the rate of ethanolamine degradation through the oxazolidone mechanism. FIGURE 2
demonstrates this difference in degradation product accumulation between JEFFTREAT® M-510
solvent and a DEA/MDEA formulation under controlled laboratory conditions.

48th Annual Laurance Reid Gas Conditioning Conference March 1-4, 1998
C:\GTData\Gas Treating Literature\Jefftreat Tech Notes\ds115698.doc Page 6
FIGURE 2. Buildup of Degradation Products:
JEFFTREAT® M-510 Solvent vs. a DEA/MDEA Formulation
Accelerated Degradation Tests in a Laboratory Autoclave
70
Oxazolidones in
(DEA or M-510 Promoter) 60 DEA/MDEA Formulation
% of Parent Compound
Degradation Products,

50

40
Piperazines in
30 DEA/MDEA Formulation

20
Inert Compound in Ethylenediamines in
JEFFTREAT® M-510 Solvent DEA/MDEA Formulation
10

0
0 1 2 3 4 5 6 7
Elapsed Time, Weeks

FIGURE 3 demonstrates that JEFFTREAT® M-510 solvent also has superior stability when
compared with a MMEA/MDEA formulation. One interesting observation from this test is that the
MMEA/MDEA formulation converts less to the substituted piperazines than the DEA/MDEA
formulation does. As noted earlier, the substituted piperazine is a desired endpoint of ethanolamine
degradation. This observation suggests that MMEA, as a promoter, may result in the accumulation
of more intermediate products (diamines) than DEA would, under similar circumstances.

FIGURE 3. Buildup of Degradation Products:


JEFFTREAT® M-510 Solvent vs. an M M E A /MDEA Formulation
Accelerated Degradation Tests in a Laboratory Autoclave
70

60 Oxazolidones in
(MMEA or M-510 Promoter)

M M E A /MDEA Formulation
% of Parent Compound
Degradation Products,

50

40
Piperazines in
M M E A /MDEA Formulation
30
Ethylenediamines in
20 M M E A /MDEA Formulation

Inert Compound in
10 JEFFTREAT® M-510 Solvent

0
0 1 2 3 4 5 6 7
Elapsed Time, Weeks

Plant Data

48th Annual Laurance Reid Gas Conditioning Conference March 1-4, 1998
C:\GTData\Gas Treating Literature\Jefftreat Tech Notes\ds115698.doc Page 7
With the increasing sophistication of process simulation software, a few operating companies have
tried experimenting with their own “formulated” MDEA solvents. In a few instances, such field
experiments have been partially successful. However, an equal number of plants have found that
such trials can be costly and expensive if they are not successful. First, the operator is “on his own”,
as it is unlikely that a formulated solvent manufacturer will offer their technical expertise and
support for a product that they themselves have not fully tested or manufactured initially. Second,
without the analytical support of a competent solvent supplier, the operator is unlikely to be aware
of ongoing degradation of the solvent until operating problems arise. Such problems can include
corrosion, foaming and equipment fouling, and potentially a decrease of capacity. Once this
situation has occurred, it is often too late to salvage the solvent.

The laboratory study of the DEA/MDEA blend showed THEED converted into bis-HEP to the
extent that, at the end of the test, approximately equal amounts of each were present. In actual
operations however, samples from numerous plants indicate that this last step in degradation does
not appear to be easily achievable. It is our belief that the substituted ethylenediamine undergoes a
dehydration process to form the corresponding piperazine via acid catalyzation. The dehydration
process in an aqueous amine solution generally requires higher temperatures than normally
encountered in plant operations.

Table II shows analytical results for several DEA and DEA/MDEA plants. Note that in each case
presented, THEED is present in greater concentration than bis-HEP. Additionally, in many cases
only traces of bis-HEP were detected.

Table II. Analysis of Degraded Solutions from Operating Plants


DEA/MDEA Blends DEA Only
Plant 8a 8b 9 10 11 13 14 15 16 17
DEA, % Weight 10.4 10.1 14.9 13.4 17.7 20.7 16.4 20.5 20.9 28.3
THEED, % Weight 4.5 3.0 4.4 9.1 6.5 9.4 9.6 5.1 1.7 4.6
Bis-HEP, % Weight 1.8 1.3 Trac 0.5 0.2 2.3 2.3 trace 0.3 trace
e
“chained-out” EA’s +* ND + + + + ND ND ND
Bis-HEP/THEED ratio 0.4 0.4 -- 0.05 0.03 0.2 0.2 -- 0.2 --

+ identified in substantial concentration but not quantified


+* identified in high concentration but not quantified

Perhaps with long-term operation, more THEED would convert into bis-HEP in these plants.
However it is more likely, due to problems associated with the increase in concentration of THEED,
the degraded solvent either would be discarded and replaced with new solvent, or losses and
makeup due to upsets would minimize this final degradation step.

FIGURE 4 shows the accumulation of degradation products over time from three different plants.
In both the DEA and blended MDEA solutions, THEED was found as the dominant degradation
product, and bis-HEP was present in much lower concentration.

48th Annual Laurance Reid Gas Conditioning Conference March 1-4, 1998
C:\GTData\Gas Treating Literature\Jefftreat Tech Notes\ds115698.doc Page 8
FIGURE 4 - Amine Degradation Comparison
High-CO2, Low-H2S Gas
40
DEA
35
% Degradation of Amine

30
DEA-Blend
25

20

15

10
JEFFTREAT® M-510
5

0
0

100

200

300

400

500

600
Days from Startup w/fresh Solvent

M-510 Blend DEA

DEGRADATION EFFECTS ON OPERATIONS

Several factors are involved in assessing the effects solvent degradation can have on individual
unit operations. Corrosion and loss of capacity (and/or the need to continuously makeup
promoter, as indicated by Dawodu and Meisen) are two of the likely problems encountered when
solvent degradation occurs.

During engineering evaluation of process options, cost analysis of the equipment life and designing
with adequate capacity are two very essential issues. Morrow states that the cost for equipment
repair and/or replacing the amine solution every 6 months can be staggering. Morrow also states
that by using a solvent that does not exhibit corrosive degradation characteristics, installed plant cost
can be reduced by 25%. Unfortunately, predicting when problems such as degradation are going to
occur is still a difficult task.

CORROSIVITY

Helle (1995) presents a mechanism for the corrosive nature of amine solutions, which is based on
the presence of complexing or chelating diamines. McCullough and Nielsen (1996) also state
that diamines can complex iron, nickel, and chromium ions. Helle says that complexants promote
depassivation, and interfere with repassivation. According to Helle, in the presence of chelating
agents, the dissolved iron does not participate in a passivating redeposition of iron-carbonate,
iron-sulfide or iron-oxide. Iron remains dissolved in the degraded amine solution in the form of
various Fe2+-complexes and then the iron is carried away from the metal surface. Chelating
agents coordinate with cations to form chelates. In the presence of chelating agents, the metal
surface is therefore laid bare and open to further iron dissolution and corrosion.

48th Annual Laurance Reid Gas Conditioning Conference March 1-4, 1998
C:\GTData\Gas Treating Literature\Jefftreat Tech Notes\ds115698.doc Page 9
Corrosion testing of degraded solutions have shown higher corrosion rates and corrosion
potential than solutions with no degradation. Helle reported an aqueous solution of 22% DEA +
2.5% THEED (+ others) showed a free corrosion current density and passivating current density
that was about 1 to 2 orders of magnitude higher than an uncontaminated 30% DEA sample.
Chakma and Meisen reported measuring corrosion rates 6.1 times higher for partially degraded
DEA solution than that obtained for an undegraded solution. They also reported that degraded
DEA solutions have an increased pitting potential, further explaining the experience of numerous
field applications that have shown severe pitting of plant equipment.

It is common to find high concentration of soluble metals when degradation products of the
ethanolamines are prevalent. McCullough and Nielsen (1996) mention that degraded DEA
solutions can dissolve up to 6000 ppmw iron (ferric ion), and that degradation products can
sometimes attack stainless steel as fast as carbon steel. The ability of degraded ethanolamines to
dissolve metals is particularly pronounced in applications in which the ratio of CO2 to H2S is high
(e.g., greater than 10). Increased solubility of metals can be a factor in operating difficulties such as
foaming, fouling and frequent filter changes. It can also be an indicator of a potentially corrosive
situation.

In operating plants that use DEA or formulated MDEA solvents that contain ethanolamines, detailed
solution analysis has shown that as degradation proceeds, the concentration of soluble metals in the
amine solution increases. Figure 5 shows the soluble metals content of the three plants referenced
in Figure 4. In the DEA-containing plants, as the level of degradation increases over time, the
concentration of soluble metals also increases. However, the relationship appears to be non-linear
in this set of data.

FIGURE 5: Soluble Metals Comparison


H igh-CO2, Low-H2S Gas
400

350
Fe - Blend
300
Fe or Cr, ppmw

250

200
Fe - DEA Cr - Blend
150
Fe - M-510
100 Cr - M -510

50

0
0

100

200

300

400

500

600

Days from Startup w/fresh Solvent

In treating coal seam gas and other gas streams containing CO2 as the only acid gas, several
operators/contractors have instituted strict guidelines for piping and equipment metallurgy. With

48th Annual Laurance Reid Gas Conditioning Conference March 1-4, 1998
C:\GTData\Gas Treating Literature\Jefftreat Tech Notes\ds115698.doc Page 10
the use of JEFFTREAT® M-510, it is possible to reduce these strict metallurgy requirements and
save a substantial amount of money for capital and equipment replacement cost. Certain
requirements are still prudent to minimize the corrosion aspects of handling wet CO2.

FIGURE 6 shows spot sample results from 17 different plants. Degraded solutions tend to have
higher concentration of soluble iron. This result is consistent with the observation that THEED
and similar degradation products can complex iron.
FIGURE 6 - Soluble Iron and Degradation
Sam p les from L o w - or No-H2S Plants
500 75
450
400

Excess Amine, % of Amine


Soluble Iron, ppmw

350 Degradation
50
300
Online conversion from
250 ethanolamine blend
200
Soluble Iron
25
150
100
50
0 0
1

8a

8b

10

11

12

13

14

15

16

17
JEFFTREAT® M-510 Plants DEA Blends DEA

Soluble Iron Excess Amine

ALKALINITY DOES NOT REVEAL DEGRADATION

Substituted ethylenediamine compounds still contain basic nitrogen atoms, and in the case of
THEED and DMHEED, only one of the nitrogen atoms is a secondary amine. This suggests that the
degradation product will be reactive towards CO2, but potentially not as reactive as the two moles of
the parent amine.

McCullough and Nielsen make the point that in MEA degradation, the HEED which results has two
basic nitrogen amines (pka1 = 10.1, pKa2 = 7.2). Since alkalinity titration typically uses an
endpoint in the range of pH = 5 - 6, it is likely that both nitrogen atoms in HEED will be registered
in the titration. In the case of THEED and DMHEED, both nitrogen atoms are likely to titrate as
amine in field alkalinity tests. For substituted diethylenetriamines, it is likely that two of the nitrogen
molecules will titrate successfully.

It is possible to have a severely degraded solution of ethanolamine and be unaware of the problem.
The following solution analysis (Table III) is from an application utilizing DEA to remove CO2
from natural gas. This plant had been in service for only six months, and the operator was unaware
of the degradation of the amine:

48th Annual Laurance Reid Gas Conditioning Conference March 1-4, 1998
C:\GTData\Gas Treating Literature\Jefftreat Tech Notes\ds115698.doc Page 11
Table III. Solution Analysis from a DEA/CO2 Plant
DEA, % Weight 18.7 In this example, more than
THEED, % Weight 7.5 25% of the amine was
Total, % Weight 26.2 converted to THEED in only
six months of operation.
This occurred without an
Alkalinity, % Weight as DEA 27.9 apparent change in alkalinity.

Without a detailed analysis and complete material balance on the solvent, including identification
and quantification of the amine itself, the build-up of degradation products can go completely
unnoticed until a problem occurs. Huntsman Corporation’s gas treating analysis program is one of
the most extensive in the industry. Our routine analysis program does several basic material
balances and if unacceptable values are found, then more extensive testing is provided.

Attachment 1 shows the results of a sample analysis for a DEA plant, (name and address withheld)
that has experienced extreme corrosion and operating problems. This plant is less than a year old
and has already found severe pitting in the reboiler tubes. Unfortunately for the plant, routine
samples had not been sent in and the degradation was not found until problems began to occur.
Note that, if only basic testing had been done on this solvent (e.g., acid gas loadings, alkalinity,
foaming and heat stable salt content), no problems with the solvent would have been detected. With
iron at 120 ppm and chromium at 409 ppm, the solution is definitely in a corrosive state. Metals
reported in this example are “soluble metals”1 not total metals. This allows for the inference that the
metals are held in solution by objectionable contaminants (chelates) derived from the degradation of
the parent amine and are not present simply due to poor filtration efficiency.

By conducting a detailed analysis and making material balances, the sample analysis program can
clearly show that “other” amines are present which would otherwise go undetected. The amine
balance block clearly shows that there is 4.4% excess amine in this sample. Subsequent analysis
using NMR confirmed the presence of 4%w of THEED. Without such a rigorous sampling
program, plant operators lack all the information they need to make informative decisions on how to
best solve their problems.

SUGGESTIONS FOR MINIMIZING DEGRADATION

When a degradation problem is found, the most common question is “what can be done to stop
the problem?” In using ethanolamines or ethanolamine-containing formulations, some design
and/or operating changes can be made to help decrease the rate of degradation, but not eliminate
it. Additionally, once the degradation has occurred, it cannot be reversed. Reclaiming of the
material may help improve solvent performance, depending on the style of reclaiming employed.
However, reclaiming is unlikely to remove some of the degradation products, and it may result in
changes in solvent formulation. Considering these factors, operators can help minimize the
degradation by monitoring and maintaining the following parameters:

1. Lower skin temperatures – Temperature has a dramatic influence on the level of


degradation. Proper design of the reboiler to reduce the heat flux and lowering the tube
skin temperature will greatly help.

1
meaning, analyzed after filtration with a small micron filter.

48th Annual Laurance Reid Gas Conditioning Conference March 1-4, 1998
C:\GTData\Gas Treating Literature\Jefftreat Tech Notes\ds115698.doc Page 12
2. Adequate stripping – If the regeneration process is operated such that higher levels of
CO2 are present in the lower section of the tower, the solvent is subjected to high
temperatures when more CO2 is present. Although the lean loading may not indicate
excessive CO2, if the major portion of the regeneration is occurring in the bottom section
of the tower or in the reboiler itself, degradation will be accelerated.

3. Lower rich loadings – Although many of the amines have higher capacity from an
equilibrium standpoint, maintaining a lower rich amine loading will help decrease the
product degradation. Lower rich loadings improve the CO2 distribution profile in the
regenerator, thereby reducing the CO2 content in the hottest portion of the regeneration
process.

4. Amine concentration – Operating with lower amine concentration decreases the boiling
point of the amine solution and increases the stripping efficiency. Although the skin
temperature is much more critical, the boiling point of the solution will also play a role in
the rate of degradation.

5. Stripper pressure -- In addition to amine concentration, stripper pressure also controls


the boiling point of the solution. Therefore operating at a minimum practical stripper
pressure will help to minimize degradation.

6. Adequate monitoring of solvent quality – Adequate analytical support is necessary to


understand the composition of your solvent and monitor it as it degrades. With results
indicating solvent degradation, operating parameters can be changed in an attempt to
minimize degradation. Huntsman Corporation’s extensive testing of all of Huntsman’s
gas treating products is one of the primary reasons the data presented here is available.

7. Chemical Selection - The type of amine solvent will greatly effect the degradation
potential of the system. Utilizing a product such as Huntsman Corporation’s
JEFFTREAT® M-510 solvent that is specifically formulated to avoid ethanolamine
degradation is a major step in improved plant operations and reducing corrosion
problems.

CONCLUSIONS

Numerous studies have shown that degradation occurs when ethanolamines are used to treat gas
that contains CO2. Recent evidence has highlighted the likelihood that some of these degradation
products complex metals, and therefore may increase the corrosivity of degraded solutions.
Although degradation does not occur in every plant, the consequences of degradation and
corrosion can be catastrophic in terms of amine consumption and equipment replacement.
Utilizing proper operating procedures can reduce the degree of degradation but not eliminate it.
Proper selection of amine solution can greatly reduce and/or eliminate the problems associated
with ethanolamine degradation.

Laboratory results demonstrated that DEA and MMEA are subject to degradation when
formulated with MDEA and when exposed to CO2 at stripper temperature. The rates of
degradation were similar between DEA and MMEA. In contrast, JEFFTREAT® M-510
demonstrated comparative stability under these conditions.

48th Annual Laurance Reid Gas Conditioning Conference March 1-4, 1998
C:\GTData\Gas Treating Literature\Jefftreat Tech Notes\ds115698.doc Page 13
Plant sample results demonstrated that DEA degradation in DEA/MDEA formulations is a
problem in actual practice. In these samples, THEED is present in higher concentration than bis-
HEP.

Sample history was used to demonstrate that, as degradation proceeds, the concentration of
metals (iron, in particular) increases in solution. The possibility that ethyleneamine compounds
can complex iron was discussed as an explanation for the increase in metals concentration. Plant
sample results also demonstrated the stability of JEFFTREAT® M-510 in this service and lack of
the iron buildup in the solvent over time.

REFERENCES

Blanc, C. M. Grall and G. Demarais, “The Part Played by Degradation Compounds in the
Corrosion of Gas Sweetening Plants Using DEA and MDEA”, Presented 1982 Gas Conditioning
Conference, Norman, OK.

Chakma, A. and A. Meisen, “Corrosivity of Diethanolamine Solutions and Their Degradation


Products”, Ind. Eng. Chem. Prod. Res. Dev., Vol. 25, No. 4, 1986.

Chakma, A. and A. Meisen, “Degradation of Aqueous DEA Solutions in a Heat Transfer Tube”,
The Canadian Journal of Chemical Engineering, Vol. 65, April 1987.

Dawodu, O.F. and A. Meisen, “Degradation of Alkanolamine Blends by Carbon Dioxide”, The
Canadian Journal of Chemical Engineering, Vol. 74, December 1996.

Helle, H.P.E., “Guideline for Corrosion Control in Alkanolamine Gas Treating, New Plantation
bv., Delft, Holland, Second Addition, January, 1995.

Hsu, C. S and C. J. Kim, “Diethanolamine (DEA) Degradation under Gas-Treating Conditions”,


Industrial Engineering Chemistry, Prod. Res. Dev., Vol. 24, No. 4, pp. 630-635, 1985.

Kim, C. J. and G. Sartori, “Kinetics and Mechanism of Diethanolamine Degradation in Aqueous


Solutions Containing Carbon Dioxide”, International Journal of Chemical Kinetics, Vol. 16, pp.
1257-1266, 1984.

Kohl, A. and R. Nielsen, “Gas Purification,” 5th Ed., Gulf Publishing, Houston, 1997.

Lang, F. S. and J.F. Mason, Jr., “Corrosion in Amine Gas Treating Solutions”, Presented at the
Thirteenth Annual Conference, National Association of Corrosion Engineers, St. Louis,
Missouri, March 11-15, 1957.

McCullough, J. G. and R. B. Nielsen, “Contamination and Purification of Alkaline Gas Treating


Solutions”, Corrosion96, Paper No. 396, 1996.

Morrow, D., “New Solvent Minimizes Corrosion in Coal Seam Gas Amine Plants”, Presented at
Rocky Mountain GPA Meeting, September 1997.

48th Annual Laurance Reid Gas Conditioning Conference March 1-4, 1998
C:\GTData\Gas Treating Literature\Jefftreat Tech Notes\ds115698.doc Page 14
ATTACHMENT 1
HUNTSMAN Gas Treating Sample Results 971100.01
Manufacturers of JEFFTREAT® Products and DGA® Agent for your gas treating needs.
Customer Challenge Phone: 123-456-7890
Plant with Problems Fax: 098-765-4321
Amine: DEA Sample #: 971100.01
Type: LEAN Received: 11/26/97
Application: High CO2/ H2S Gas Plant Reported: 12/09/97
Unit ID: Plant Other ID: 11/17/97
BULK COMPOSITION, %w
Usable Amine 31.95 High
Water + 66.40 OK Alternate Units
H2 S + 0.00 OK 0.0 gr/gal 0.0000 m/m
CO2 + 0.35 OK 0.26 scf/gal 0.0259 m/m
+
Total Loading = 0.0259 m/m OK
Neutralization Calculations %w Amine Basis
Strong Acid Anions + 0.02 OK Speciated Acids: 0.04 Avg.
nFDEA + < 0.10 OK Titrated Acids: 0.14 0.08 OK
Strong Bases + 0.00 OK Strong Bases: 0.01
Bound Amine + 0.08 OK Acids-Bases: = 0.08
% Neutralization = 8 OK

Sample Recovery = 98.90 OK Additional Comments:


NITROGEN BALANCE AMINE BALANCE
%w as Amine
Total N: - 32.6 %w as Amine
In Titr. Am: - 32.0 Titrated: . 32.0
In Anions: - 0.0 Bound: + 0.1
In Bound Am: - 0.1 Identified: - 27.6
Excess N: = 0.05 OK Excess Am: = 4.4 High
SOLUBLE METALS ANION SPECIATION
ppmw ORGANIC ppmw INORGANIC pppw
Iron 120 High Acetate <100 OK Chloride 31 OK
Chromium 409 High Formate + <100 OK
Nickel <1 OK Glycolate + <100 OK Sulfite + <200 OK
Manganese <0.5 OK Lactate + 155 OK Sulfate + trace OK
Sodium 16 Oxalate + trace OK Thiosulfate + <100 OK
Potassium <5 Propionate + <100 OK Thiocyanate + <100 OK
Calcium <1 OK Total * = 161 OK Total* Inorg. = 77 OK
Silicon <10 Total* speciated Anions = 238 *Totals include traces
FOAMING TENDENCY PHYSICAL PROPERTIES
Rate Volume, cc Color: Dark Green / Haze Caution
250 61 nil For 32.0 %w amine Expected Actual
250 61 nil Breaktime, s DENSITY g/ml 60°F: 1.0435 1.0438 OK
500 70 Nil 7 VISCOSITY Cs 100°F: 2.31 2.41 OK
This solvent contains <50 ppmw ammonia. OK

48th Annual Laurance Reid Gas Conditioning Conference March 1-4, 1998
C:\GTData\Gas Treating Literature\Jefftreat Tech Notes\ds115698.doc Page 15
FOR MORE INFORMATION OR LITERATURE
Please call the nearest Huntsman Corporation office.

HUNTSMAN CORPORATION Trading Corporation Woluwe Office Garden


3040 Post Oak Boulevard 350 Orchard Road Woluwedal 26
Houston, TX 77056 #11-07 Shaw House B-1932 Zaventem, Belgium
Tel: 713-235-6000 Singapore 238868 Tel: 32-2-718-0120
Fax: 713-235-6977 Tel: (65) 730-0270 Fax: 32-2-718-0211
Fax: (65) 730-0280
Huntsman Corporation Canada Inc.
Research and Development Huntsman de Brasil Partcipacoes Ltda 256 Victoria Road,South
7114 North Lamar Boulevard R. Helosia Pamplona, 628 Guelph, Ontario N1E 5R1, Canada
Austin, TX 78752 Sao Caetano Do Sul - SP 09520 Tel: 519-824-3280
Tel: 512-459-6543 Sao Paulo, Brazil Fax: 519-824-4979
Fax: 512-483-0925 Tel: 55-11- 442-9264
Fax: 55-11-441-7216 Huntsman de Mexico, S.A. de C.V.
Rio San Joaquin #305
REGIONAL CUSTOMER Huntsman Corporation C.A. Colonia Pensil
SERVICE REPRESENTATIVES Multicentro Paseo El Parral Mexico City, Mexico 11480
Piso 6, Oficina 11 Tel: (52) 5-280-1476
(Las Cuatro Avenidas) Urb. El Parral 5-280-1405
Midwest Valencia Estado Carabobo 5-280-1485
Tel: 1-800-231-3107 Venezuela Fax: (52) 5-280-1494
Tel: 58-41-25-4547
Southeast Fax: 58-41-25-2267
Tel: 1-800-624-6419 Emergency Assistance
For transportation emergencies only,
Northeast call CHEMTREC 1-800-424-9300.
Tel: 1-800-231-3104

North Central For all other emergencies, call


409-722-8381, our 24-hour emergency
Tel: 1-800-624-6417
number in Port Neches, Texas.
West Coast/Texas
Tel: 1-800-826-0868

Huntsman Corporation Europe


Huntsman International Huntsman Corporation Belgium N.V.
Copyright  1998 HUNTSMAN CORPORATION
Huntsman Corporation warrants only that its products meet the specifications stated herein. Typical properties where stated, are to be considered as representative of current
production and should not be treated as specifications. While all the information presented in this document is believed to be reliable and to represent the best available data
on these products, NO GUARANTEE, WARRANTY, OR REPRESENTATION IS MADE, INTENDED, OR IMPLIED AS TO THE CORRECTNESS OR SUFFICIENCY OF ANY
INFORMATION, OR AS TO THE SUITABILITY OF ANY CHEMICAL COMPOUNDS FOR ANY PARTICULAR USE, OR THAT ANY CHEMICAL COMPOUNDS OR USE
THEREOF ARE NOT SUBJECT TO A CLAIM BY A THIRD PARTY FOR INFRINGEMENT OF ANY PATENT OR OTHER INTELLECTUAL PROPERTY RIGHT. EACH USER
SHOULD CONDUCT A SUFFICIENT INVESTIGATION TO ESTABLISH THE SUITABILITY OF ANY PRODUCT FOR ITS INTENDED USE. Products may be toxic and require
special precautions in handling. For all products listed, user should obtain detailed information on toxicity, together with proper shipping, handling, and storage procedures,
and comply with all applicable safety and environmental standards.
Main Offices: Huntsman Corporation / P.O. Box 27707 / Houston, Texas 77227-7707 / (713) 235-6000
DS115698 (6/3/98) Technical Services Section: P.O. Box 15730 / Austin, Texas 78761 / (512) 483-0056

Anda mungkin juga menyukai