You are on page 1of 50

9

NEUTRALIZING AND FILMING AMINE TREATMENT

9.1 Introduction to Amine Treatments


Amines are a large class of organic compounds containing a nitrogen group structurally
resembling ammonia (NH3) in which one or more hydrogen atoms are replaced with an organic
radical (e.g. NRH2, R2NH, or R3N) [1]. Each hydrogen ion in ammonia (NH3) can be substituted
with an organic compound, e.g., methyl (CH3), ethyl (C2H5), propyl (C3H7), etc. to give a series
of compounds with one to three carbon chains of increasing lengths. Either a single hydrogen of
ammonia can be substituted, giving CH3NH2 (methyl amine), C2H5NH2 (ethyl amine), C3H7NH2
(propyl amine), etc.: two hydrogen ions can be substituted, giving (CH3)2 NH (di-methyl amine),
(C2H5)2 NH (di-ethyl amine), (C3H7)2 NH (di-propyl amine), etc.: or all three hydrogen ions can
be substituted, giving (CH3)3N (tri-methyl amine), (C2H5)3N (tri-ethyl amine), (C3H7)3N (tri-
propyl amine), etc.
It is possible to have two (or more) amine groups in the same compound (polyamines).
Examples include diamines, e.g., diaminoethane (NH2C2H4NH2), diaminopropane
(NH2C3H6NH2) and triamines, e.g., triaminopropane (NH2C3H5(NH2) 2), with an amine group on
each carbon atom.
In the combined cycle / HRSG plant, amines that may be applied for chemistry and corrosion
control purposes come under two broad categories:
 Neutralizing Amines: these are amines which are primarily applied to raise the pH of
condensate and feedwater system of the plant, and
 Filming Amines (or film forming amines): these are amines that adhere to metallic surfaces
of the plant cycle to form a monomolecular “film” or barrier.
In some applications neutralizing and filming amines are used together in a blend.
The neutralizing amines behave in much the same way as ammonia, reducing the solubility of
metallic oxides formed by raising the pH of the condensate, feedwater or evaporator/drum water.
Filming amines provide corrosion protection by forming a physicochemical barrier between the
metallic surfaces and the working fluid (water) to prevent corrosion from occurring. Filming
amines also can provide a film on the steam surfaces and offer protection against oxygen pitting
when units are off line and exposed to humidified air or water formed via condensation.
The properties of some of the neutralizing amines have the potential to improve the pH
conditions in the low pressure (LP) and intermediate pressure (IP) evaporators and economizers
of HRSGs, the phase transition zone (PTZ) of the LP steam turbine, the condensing steam in wet
and air cooled condensers (WCC and ACC), and the pH conditions at any other two phase flow
location. The potential to improve pH in these environments arise from lower volatility to
improve distribution of the amine to the liquid phase in low pressure applications and higher

9-1
Neutralizing and Filming Amine Treatment

basicity due to higher dissociation of the amine, resulting in higher pH at operating temperatures.
EPRI has recently produced several reports investigating the use of neutralizing amines in fossil
plants for these purposes; these reports are the foundation for this Chapter [1-3].
Film forming amines have properties to provide a corrosion inhibiting barrier to oxygen, carbon
dioxide, and moisture that can supplement layup practices and perhaps offer a viable alternative
for corrosion protection in troublesome environments of some operating units. In addition, there
are some fossil plants where these products have been used on a continual basis [4-11]. The
strong interaction of the filming amine and the metal surface is a result of chemisorption between
the polar amino nitrogen group and the unfilled orbital of iron (Fe) atoms on the surface of the
metal. The hydrophilic end of the filming amine attaches to the metal surface and the filming
amine’s hydrophobic end, which repels water, faces outward. In 2009/2010, EPRI commissioned
a field assessment of equipment layup protection with the generic filming amine octadecylamine
(ODA), C18H37NH2, to make a scientific assessment of the advantages and disadvantages of using
a filming amine (mainly ODA) to protect idle equipment of the steam-water cycle in fossil units
[1]. This study examined the practices and observation of eight (8) power plants totaling 25
units, including drum, once-through and supercritical units. The determination of the study
concluded that while ODA provided good protection, the disadvantages of the product including
immiscibility, solidification, iron sloughage, and fouling tendencies were unacceptable. In 2010,
EPRI initiated a detailed study on the use of a proprietary filming amine displaying less
undesirable properties for shutdown protection. At the time of publishing this research is ongoing
and results are not available for publishing, however a wealth of experience reports have been
issued both by utilities participating in and outside of this research effort [4-11].

9.2 Neutralizing Amines Background Science


This Chapter discusses six neutralizing amines summarized in Table 9-1.
Table 9-1
Neutralizing Amines

Chemical Molecular
Amine Abbreviation Structure
Formula Weight (g/mol)

Dimethylamine DMA C2H7N 45.08

Ethanolamine ETA* C2H7NO 61.08

5-Aminopentanol 5AP C5H13NO 103.16

3-Methoxypropylamine MPA C4H11NO 89.14

Morpholine Morph C4H9NO 87.1

Cyclohexylamine CYC C6H11NH2 99.2

*Ethanolamine is the proper chemical name of Monoethanolamine (MEA) which is frequently the chemical name
used in industry, similarly 3-Methoxypropylamine (MPA) is the proper chemical name of MOPA.

9-2
Neutralizing and Filming Amine Treatment

When discussing the potential application of neutralizing amines there are three properties that
are critical to understand in order to determine whether the amine will provide a benefit in an
application:
1. Dissociation – This is the relative ionization of the amine, for ammonia this is essentially the
degree to which dissolved ammonia gas ionizes in water:
NH3(aq) + H2O NH4+ + OH-
For dimethylamine the ionization in water is:
H3C H3C
+
NH + H2O NH2 + OH-
H3C H3C
For ethanolamine the ionization in water is:
HO-C2H4-NH2(aq) + H2O HO-C2H4-NH3+ + OH-
For 5-aminopentanol the ionization in water is:
HO-C5H10-NH2(aq) + H2O HO-C5H10-NH3+ + OH-
For 3-methoxypropylamine the ionization in water is:
H3C-O-C3H6-NH2(aq) + H2O H3C-O-C3H6-NH3+ + OH-
For morpholine the ionization in water is:
+
H-N H2-N
+ H2O + OH-
O O

For cyclohexylamine the ionization in water is:


NH2 NH3+

+ H2O + OH-

2. Distribution – This is the relative distribution of the amine between a gas and liquid phase in
a two phase environment, for ammonia this is essentially the degree to which aqueous
ammonia gas is liberated to steam in a two phase environment.
3. Decomposition – This is the relative amount of amine that thermally decomposes in the
cycle, typically amines decompose into ammonia, organic compounds (e.g. organic acids,
non-ionic organics) and carbon dioxide.
Each of these properties varies with temperature and pressure.

9-3
Neutralizing and Filming Amine Treatment

9.2.1 Estimating Iron Solubility Impact of Applying Neutralizing Amines


The key aspect for understanding the potential benefits of neutralizing amines is to determine
their expected results on iron solubility. Neutralizing amines have potential benefit over
ammonia primarily in two phase conditions where higher liquid phase pH is possible than with
ammonia alone. In two phase conditions, the liquid phase will be deoxygenated due to oxygen’s
high partitioning coefficient. Since the liquid phase is deoxygenated, iron solubility will follow
the solubility of magnetite given in Figure 2-9 which is reproduced here as Figure 9-1. It is well
established that at-temperature pH has a significant effect on flow-accelerated corrosion (FAC)
rates through its impact on iron solubility [12-16]. Figure 9-1 illustrates the effect of temperature
on iron solubility of ammonia solutions at different concentrations. This implicitly shows the at-
temperature pH (pHT) effect on iron (magnetite) solubility across different temperatures since
ammonia affects iron solubility by altering pH. This figure can then be used with the known at-
temperature pH of ammonia to provide a method to estimate iron solubility improvements when
applying various neutralizing amines through their effect on at-temperature pH.
As seen in Figure 9-1 iron solubility is minimized across the temperature range, if the ammonia
concentration is raised enough to attain a 9.4 to 9.6 pH at 25ºC (77ºF). The at-temperature pH
influence on iron solubility should be the same regardless of the neutralizing agent applied. If a
neutralizing amine can be applied so that the at-temperature pH is the same or greater than the at-
temperature pH of a single phase ammonia solution with a concentration sufficient to attain 9.4
to 9.6 pH at 25ºC (77ºF), then presumably this amine will result in similar temperature dependent
iron solubility as shown in Figure 9-1.
As mentioned earlier, Figure 9-2 provides at-temperature pH of ammonia under single phase
conditions, i.e. it only considers the first ‘D’, dissociation. Ammonia solutions under two phase
conditions will not attain these at-temperature pH values in the liquid phase due to the second
‘D’, distribution. Ammonia will preferentially distribute into the steam phase reducing the liquid
water phase concentration and pH (both at-temperature and at 25ºC). This will result in elevated
iron solubility in the liquid phase under two phase condition. If the pH of the liquid phase is
known then Figure 9-1 can be used to estimate this increase in iron solubility. Software
applications such as EPRI’s MULTEQ, multi-phase equilibrium program, can estimate this final
pH at-temperature in the liquid phase whether ammonia or another neutralizing amine is being
applied under two phase conditions.
By having the at-temperature pH, at the temperature simulated, one can find the corresponding
ammonia concentration that would yield that at-temperature pH at the given temperature (in
Figure 9-2). One can then establish the corresponding pH at 25ºC (77ºF) and from that, infer,
using Figure 9-1, what the iron solubility would be at the given temperature in the liquid phase of
the two phase conditions. In this way one can estimate the potential relative improvement on iron
solubility applying various neutralizing amines under various HRSG steam / water cycle two
phase conditions. This methodology does not purport to determine actual iron solubility results,
but does provide a common basis for comparison and is used throughout this chapter to assess
relative improvements by applying various neutralizing amines.

9-4
Neutralizing and Filming Amine Treatment

Figure 9-1
Solubility of Magnetite as a Function of Temperature at Various Ammonia Concentrations
Adapter from Reference [17]

9-5
Neutralizing and Filming Amine Treatment




 






      

 

Figure 9-2
Ammonia in Water pHT over Fossil Power Plant Water Cycle Temperatures [18]

Combining Figure 9-1 and 9-2 allows one to map a given pHT back to iron solubility based on
the equivalent ammonia concentration needed to achieve the pHT (Figure 9-3).

Figure 9-3
Mapping At-Temperature pH and Corresponding Iron (Magnetite) Solubility Curves

9-6
Neutralizing and Filming Amine Treatment

9.2.2 Dissociation of Neutralizing Amines


Figure 9-4 gives the effect of temperature on the dissociation constant (given by pKb) for a
variety of amines. Note the lower the pKb the stronger the base, so DMA is the strongest base
and depending on the temperature ammonia (NH3) and morpholine (Morph) are the weakest.
Table 9-2 gives the concentration of various amines required to achieve a pH of 9.6 at 25ºC.

Figure 9-4
Dissociation: Effect of Temperature on pKb of Common Amines (Lower pKb = Stronger
Base)

Table 9-2
Concentration of Various Amines necessary to achieve a pH25 of 9.6

Amine ppm pH25 = 9.6 mmol/L pH25 = 9.6


Morph 48 0.551
NH3 2.3 0.134
ETA 5.5 0.090
MPA 5.4 0.061
5AP 4.7 0.046
DMA 1.9 0.042
CYC 4.4 0.044

9-7
Neutralizing and Filming Amine Treatment

Using the dissociation constants from Figure 9-4 and the starting concentrations of each of the
amines it is possible to calculate the at-temperature pH of each amine when the initial pH25 = 9.6.
Figure 9-5 illustrates the results. All the neutralizing amines starting at a pH25 = 9.6 achieve the
same or higher pHT (pH at-temperature) as ammonia at 9.6. Referring to Figure 9-1 it should be
expected that each of the neutralizing amines if applied to achieve a pH25 = 9.6 will achieve the
same or better iron solubility as ammonia across the combined cycle temperature range.

 

 

 

 





      

 

Figure 9-5
Amines at pH25 = 9.6 plotted with Ammonia in Water pHT over Power Plant Water Cycle
Temperatures

From Figure 9-5 neutralizing amines provide superior or equivalent at-temperature pH to


ammonia across the temperature range in a HRSG plant under single phase conditions based on
their dissociation. However the improvement is only marginal when looking just at single phase
conditions.

9.2.3 Distribution of Neutralizing Amines


Figure 9-6 gives the effect of temperature on the distribution constant (given by Log Kd) for a
variety of amines. With distribution constants in combination with dissociation constants it is
possible to evaluate an amines impact on liquid phase pH under two phase conditions. Note the
lower the Log Kd the stronger the affinity for the liquid phase, so 5AP and ETA have the
strongest affinity for the liquid phase and DMA (followed by ammonia, NH3) have the least
affinity for the liquid phase. Also note that Kd is the ratio of the molar concentration in the steam

9-8
Neutralizing and Filming Amine Treatment

phase divided by the un-ionized molar concentration in the water phase (as such the degree of
ionization (dissociation) will influence amine volatility, in general detailed iterative calculations
must be performed to solve actual distribution of amine or ammonia in water and steam under
two phase conditions using a software package such as EPRI’s MULTEQ).

Figure 9-6
Distribution: Effect of Temperature on Log Kd of Common Amines (Lower Log Kd =
Stronger Affinity for the Liquid Phase)

Distribution is relevant in two phase environments at temperatures less than 300ºC; above this
temperature iron solubility is negligible provided the pH is above pH 7 and as such distribution
at these temperatures does not have a significant effect on general iron corrosion (or FAC). The
condenser, and the LP, IP and HP evaporators are the two phase environments in the typical
combined cycle plant. Reviewing Figure 9-1, while the HP evaporator does have a two phase
environment it operates at a temperature above which there should be appreciable iron solubility.
When evaluating amines versus ammonia the main objective is to examine conditions in
condensers and LP and IP evaporators. Figures 9-7 to 9-9 indicate the relative performance of
different amines against iron solubility in terms of the achieved at-temperature pH using the
methodology established at the beginning of this section. For this evaluation the assumed
condenser operating temperature was 50ºC (122ºF), for the LP evaporator the temperature was
150ºC (302ºF) and for the IP evaporator the temperature was 250ºC (482ºF). For each of the
amines examined as was previously the case a concentration sufficient to achieve a pH of 9.6 at
25ºC (77ºF) was utilized. These figures are an idealized construction to illustrate whether a
potential benefit can be achieved through the use of neutralizing amines. They do not reflect any
actual application as in actual applications breakdown products would affect the achieved at-
temperature pH.

9-9
Neutralizing and Filming Amine Treatment

Figure 9-7
Condenser Condition: Equilibrium Liquid phase at-temperature pH achieved at 50ºC
versus Steam Fraction / Mass Percent Steam (Ammonia or Neutralizing Amine to pH25 =
9.6) – Simulation of Idealized Situation with no Breakdown Products in Actual Application
Breakdown Products cannot be ignored

The important range of consideration within a condenser is the high steam fraction situation (i.e.
the early condensate). Examining this region (green region of Figure 9-7), each of the
neutralizing amines achieves superior results to ammonia. In the high steam fraction range of the
two phase conditions (i.e. early condensate), the liquid phase ammonia concentration will be
around 9.0 at 25ºC (which is equivalent to pHT of ~8.3, for ammonia, at 50ºC). This is the
expected early condensate pH for steam that when fully condensed would have a pH of 9.6. This
can be mapped back to Figure 9-1 to give an expected iron solubility in the early condensate. The
drop in pH of the early condensate (in comparison to the fully condensed sample) corresponds to
a significant increase in iron solubility. For the fully condensed sample the iron solubility
corresponds to the pH 9.6 line of Figure 9-1 and the early condensate iron solubility corresponds
to the pH 9.0 line of Figure 9-1. At 50ºC the iron solubility in the early condensate thus would be
nearly 20 ppb, versus < 5 ppb for the steam when fully condensed. ETA and 5AP result in the
highest pH of the liquid phase (early condensate) under the two phase conditions due to their
much higher affinity to the liquid phase.
The red line provides a relative indication of the at-temperature pH of an ammonia solution of
sufficient concentration to achieve pH 9.4 at 25ºC. This pH corresponds to the second lowest line
on the iron solubility graph given in Figure 9-1. If the at-temperature pH given in Figure 9-7 is
greater than this line than the expected iron solubility would be lower and the potential for FAC
lower. If the at-temperature pH is less than this line than the expected iron solubility would be
higher and the potential for FAC higher. The further above this line the at-temperature pH is the
lower the expected iron solubility would be, and the further below this line the at-temperature pH
is the higher the expected iron solubility would be compared to the values given for ammonia at
pH 9.4 at 25ºC given in Figure 9-1.

9-10
Neutralizing and Filming Amine Treatment

Figure 9-8
LP Evaporator Condition: Equilibrium Liquid phase at-temperature pH achieved at 150ºC
versus Steam Fraction / Mass Percent Steam (Ammonia or Neutralizing Amine to pH25 =
9.6) – Simulation of Idealized Situation with no Breakdown Products in Actual Application
Breakdown Products cannot be ignored

In evaluating the potential improvement in an LP evaporator (Figure 9-8) a different approach is


taken than for the case of the condenser. In the condenser (Figure 9-7) the graph represented a
simulation of the dynamic two phase liquid pH conditions through the condenser as the steam is
condensed to water. In the LP evaporator this approach cannot be taken directly. This is due to
the evaporator being essentially a circulating water volume in equilibrium with a steam phase
being removed (although in the FFLP the water volume is constantly also being replaced). This
is different from the condenser which is essentially a once-through component. All steam
entering the condenser, exits as condensate in a single pass. Thus all that needs to be considered
to simulate two phase conditions throughout the condenser is the operating conditions and
starting steam conditions. In the evaporator, equilibrium will be established between the bulk
water and steam based on the operating conditions in the evaporator and incoming pH of the
feedwater. This will determine the pH of the bulk water in the evaporator. It is then this water
that would exist in transitional two phase conditions within the evaporator tubing. In a SALP
operating with a volatile treatment, this bulk water pH will be significantly different than the
feedwater pH due to ammonia being preferentially removed to the steam phase, so the bulk water
in the evaporator at equilibrium has a significantly lower pH (see Table 3-8). It is this bulk water
condition that is represented by the right side green area on Figure 9-8. The actual two phase
conditions in the evaporator tubing is not being represented (but can be presumed to be an
amplification of the results in the right side green area). Similarly for an FFLP the left side green
area in Figure 9-8 represents bulk water pH conditions rather than actual two phase conditions in
the evaporator tubing (see Table 3-8 for the results of additional calculations for ammonia in an
FFLP). Again, for the FFLP, actual two phase conditions in evaporator tubing would be an
amplification of these results. As attempting to simulate the actual two phase conditions would
add complexity to the analysis and since the first simulation of the equilibrium bulk water
conditions shows the relative effect, Figure 9-8 is confined to illustrating bulk water effects.

9-11
Neutralizing and Filming Amine Treatment

In the SALP where equilibrium between the water and steam phase will be achieved the higher
steam fractions as described above is the relevant section. Here all the amines except CYC
achieve better pH than ammonia. Ammonia is stripped under these conditions so that the
effective pH for iron solubility in the water phase is around a pH of 9.1. All the amines except
CYC and DMA achieve better than an ammonia pH at 9.4 in the water phase, thus minimizing
the iron solubility. In the FFLP the actual amount of ammonia or amine stripped is dependent on
the relative amount of steam flow (see Table 3-8), regardless though there are two notable trends,
for ETA and 5AP the pH increases as the steam fraction increases. The MPA and Morph are
essentially flat across the steam fraction, DMA, CYC and ammonia decrease with steam fraction.
This is relevant for FFLP as the water from these drums is used as feedwater to the IP and HP
evaporators. As such for DMA, CYC and NH3 the feedwater will be stripped of the amine and
may not be sufficient to provide optimal iron corrosion control through the IP and HP
economizers and the associated feedwater piping.

Figure 9-9
IP Evaporator Condition: Equilibrium Liquid phase at-temperature pH achieved at 250ºC
versus Steam Fraction / Mass Percent Steam (Ammonia or Neutralizing Amine to pH25 =
9.6) – Simulation of Idealized Situation with no Breakdown Products in Actual Application
Breakdown Products cannot be ignored

For the IP evaporator the high steam fraction equilibrium, which is analogous to bulk water
conditions at equilibrium, is the important range to consider for the same reasons as discussed for
the SALP, here ETA and 5AP achieve the highest at-temperature pH.
From the preceding figures it is clear that neutralizing amines can provide enhanced pH control
in two phase conditions over what is possible with ammonia alone. However as was the case
with dissociation, the distribution is not the whole story as decomposition will also be relevant.

9-12
Neutralizing and Filming Amine Treatment

9.2.4 Decomposition of Neutralizing Amines


Decomposition testing of neutralizing amines has been performed [2,19] and have shown that
different neutralizing amines will decompose at different rates depending on the amine, the
temperature and the pressure. Generally, decomposition increases with increasing temperature,
decreasing pressure and increasing time of exposure (i.e. the rate of decomposition is greatest in
reheaters). While temperature, pressure and amine type are known, the actual time of exposure at
high temperature is typically not known. As such the actual expected decomposition rate of
amines is very much unit specific. This becomes an even more complex situation when one
considers that in practice the cycle is not a single pass but rather continuous. Thus amines
continue to breakdown with each pass, with these breakdown products in turn potentially
breaking down. Fresh neutralizing amine is being added and breakdown products are lost with
steam and water losses. The overall composition thus depends on a multitude of unit specific
conditions that will change with changes in unit operation. At steady state conditions these
conditions will result in a unit specific equilibrium of amine and residual breakdown products.
Thus the question becomes how to deal with the risk from decomposition when the final
equilibrium point based on decomposition is unknown.
Table 9-3 identifies the carbon content of the neutralizing amines, first to achieve a pH of 9.6 at
25ºC with 100% of the amine and second, what the equivalent concentration limitation would be
to meet the 200 ppb TOC diagnostic target value given in Table 3-1 for the condensate. Note this
is an experience based target value that is not rigorously based on a detailed analysis or
experimentation. As TOC is a non-specific parameter that covers organic acids, oils, plastics and
a multitude of other potential compounds, a rigorous analysis to establish a target value for TOC
is not possible, since the potential damage caused by the TOC would depend on the actual
constituents making up the TOC value. Thus while 200 ppb of TOC was chosen for the
following analysis, it does not represent a true limiting value for the application of neutralizing
amines. The limiting values for application of neutralizing amines should be evaluated based on
corrosion product transport monitoring and evaluation of corrosion control throughout the cycle
when applying the amines. These final values may very well be significantly higher than the
values in Table 9-3. However the 200 ppb TOC target value does provide a useful starting point
for analyzing the potential effectiveness of neutralizing amines at an application rate that
conforms to existing experience on acceptable TOC levels.

9-13
Neutralizing and Filming Amine Treatment

Table 9-3
Neutralizing Amines, Concentration for pH25 = 9.6 and corresponding Carbon
Concentration

Amine Abbreviation Chemical mmol/L pH25 = Carbon (ppb) ppb Amine


Formula 9.6 pH25 = 9.6 C = 200 ppb
DMA C2H7N 0.042 1,012 376

ETA C2H7NO 0.090 2,161 509

5AP C5H13NO 0.046 2,734 344

MPA C4H11NO 0.061 2,908 371

Morph C4H9NO 0.551 26,452 363

CYC C6H11NH2 0.044 3,194 276

Based on Table 9-3 it is clear that to achieve the 9.6 pH at 25ºC with many of the amines
requires a very large carbon addition to the cycle if added as a single amine. Depending on the
decomposition rates then this can result in a huge exceedance of the condensate diagnostic target
value of 200 ppb or the steam diagnostic target value of 100 ppb for TOC. Now it is not well
established how much of a particular neutralizing amine will breakdown into carbon
dioxide, organic acids or other non-polar organics, however for reference in the case of acetate
(CH3COO-) 12 ppb as carbon will result in a cation conductivity of 0.2 µS/cm, for formate
(HCOO-) only 6 ppb as carbon is necessary. Using the 200 ppb as carbon, an initial dosing target
for the neutralizing amine can be set. However it is clear that doing this would result in
insufficient neutralizing amine addition to achieve iron solubility minimization. As such,
addition of ammonia would be required to reach the ideal pH ranges for iron corrosion control if
the 200 ppb as carbon TOC diagnostic target value was used to set the initial neutralizing amine
addition rate. The question then is would such a blend achieve superior results under two phase
conditions than ammonia can alone. Reproducing the figures for the condenser, the LP
evaporator and IP evaporator this time with ammonia at pH 9.6 and neutralizing amine up to the
200 ppb as carbon one can evaluate the potential benefit. This analysis ignores decomposition
by-products which is not appropriate in actual practice but gives an indication of whether
benefits may occur at these vastly reduced addition levels. Figures 9-10 to 9-12 thus represent
best case scenarios where acidic anionic breakdown products do not accumulate. If benefits were
not found to occur based on this best case analysis, further evaluation, with organic breakdown
products factored in, would not be warranted.

9-14
Neutralizing and Filming Amine Treatment

Figure 9-10
Condenser Condition: Equilibrium Liquid phase at-temperature pH achieved at 50ºC
versus Steam Fraction / Mass Percent Steam (Ammonia at pH25 = 9.6, Neutralizing Amine =
200 ppb as Carbon) – Simulation of Idealized Situation with no Breakdown Products in
Actual Application Breakdown Products cannot be ignored

Figure 9-11
LP Evaporator Condition: Equilibrium Liquid phase at-temperature pH achieved at 150ºC
versus Steam Fraction / Mass Percent Steam (Ammonia at pH25 = 9.6, Neutralizing Amine =
200 ppb as Carbon) – Simulation of Idealized Situation with no Breakdown Products in
Actual Application Breakdown Products cannot be ignored

9-15
Neutralizing and Filming Amine Treatment

In Figure 9-10 which is a simulation of the liquid water pH at-temperature in a two phase
environment in a condenser at 50ºC with the reduced amine concentrations of Table 9-3 in
combination with ammonia (up to pH 9.6 at 25ºC) only ETA and 5AP achieved better at-
temperature pH in the liquid phase than the equivalent at-temperature pH of an ammonia solution
in a single phase environment with a 9.4 pH at 25ºC.
In Figure 9-12 the SALP evaporators also have an improvement in at-temperature pH when ETA
and 5AP are applied even at the reduced concentrations given in Table 9-3. For the FFLP
generally all chemistry combinations achieved similar results, i.e. there is minimal benefit in the
bulk water over ammonia alone. Recall though for the evaporators, unlike in the condenser, the
simulation is of bulk water conditions at equilibrium, dynamic two phase conditions will exist
though in which case the simulated results for the SALP conditions may be more applicable
within the FFLP in which case ETA and 5AP may provide significant improvement in at-
temperature pH.

Figure 9-12
IP Evaporator Condition: Equilibrium Liquid phase at-temperature pH achieved at 250ºC
versus Steam Fraction / Mass Percent Steam (Ammonia or Neutralizing Amine to pH25 =
9.6) – Simulation of Idealized Situation with no Breakdown Products in Actual Application
Breakdown Products cannot be ignored

For the IP evaporator again only the 5AP and ETA ammonia blend achieve better than the at-
temperature pH in a single phase condition by ammonia at a pH of 9.4 at 25ºC.
Thus an ammonia and neutralizing amine blend up to 200 ppb as carbon, can potentially achieve
superior pH results in two phase environments than ammonia alone can, ignoring the
decomposition products effect on the pH at-temperature. In the real case though formation of
organic acids will affect the pH in two phase conditions as they will also concentrate in the liquid
phase, as such these cannot be ignored. The next Section examines this issue.

9-16
Neutralizing and Filming Amine Treatment

9.2.5 Example of the Use of a Neutralizing Amine in a Triple Pressure HRSG


The following discussion is from a field study of a triple pressure HRSG with a FFLP [3]. It
examines whether higher condenser pH and FFLP pH can be achieved in two phase conditions
even in the presence of decomposition products such as acetate and formate. The field study unit
had an HP drum operating at 13.1 MPa (1900 psig), the IP drum at 2.9 MPa (420 psig) and the
LP Drum at 0.43 MPa (62 psig). The temperature of the main steam and reheat steam was 566ºC
(1050ºF). The HRSGs have and use duct firing to achieve full load. The steam turbine generates
200 MW with approximately 930 klbs/hour of steam flow. The neutralizing amines added were
cyclohexylamine (CYC) and ethanolamine (ETA) at the condensate pump discharge.
Table 9-4 illustrates the four steady state concentrations of ETA, CYC, ammonia, acetate,
formate, sodium, chloride and sulfate (each at a constant amine feed and unit operating
conditions). A notable point from the condensate dataset was that the organic acid steady state
concentrations (acetate and formate) were not a strong function of the steady state amine
concentrations.
Table 9-4
Steady State* Chemical, Decomposition Product and Contaminant Levels

Sample ETA CYC Ammonia Acetate Formate Sodium Chloride Sulfate


(ppb) (ppb) (ppb) (ppb) (ppb) (ppb) (ppb) (ppb)
Condensate-1 423 510 1650 61.5 17.2 17.2 1.3 1.5
Condensate-2 765 934 2210 42.0 16.9 5.3 2.2 5.0
Condensate-3 1030 1400 1260 44.5 23.2 15.4 1.0 0.6
Condensate-4 1200 1470 1140 47.1 26.0 5.2 <0.5 1.3

*Note steady state concentrations represent the combined effect of thermal breakdown, amine addition, and loss of
amine breakdown products through water, steam and air ejection, etc. at steady state operating conditions.

9.2.5.1 Simulating at-temperature pH from Neutralizing Amines and Decomposition


Products in Field Studied HRSG
MULTEQ simulation using the amine, ammonia and contaminant values in Table 9-4 can be
used to evaluate whether the neutralizing amines in the presence of acetate and formate
decomposition products was able to achieve superior at-temperature pH conditions. Table 9-5
summarizes the results of two simulations for each set of values. One with the amines, ammonia
and all contaminants present and one where the neutralizing amines and organic acids were both
removed (i.e. ammonia alone). In each of the four cases superior at-temperature pH was achieved
with the neutralizing amine (even in the presence of the organic decomposition products)
compared to ammonia alone (without the organic decomposition products). As expected the
higher neutralizing amine concentration provided the highest increase in the at-temperature pH.

9-17
Neutralizing and Filming Amine Treatment

Table 9-5
Condenser Calculated pHT using MULTEQ Amine Blend (with Acetate and Formate) and
Ammonia Only (with no Acetate and Formate)

Sample Location Temp. Neutral ETA / CYC / NH3 NH3 pHT


deg C pHT 90% Steam 90% Steam Difference
Calculated pHT Calculated pHT
Condensate-1 Condenser 45 6.70 8.65 8.45 +0.20
Condensate-2 8.94 8.52 +0.42
Condensate-3 9.03 8.43 +0.60
Condensate-4 9.05 8.37 +0.68

Table 9-6 re-examines the amine blend, this time removing the CYC to evaluate whether it was providing any
appreciable benefit to the at-temperature pH.

Table 9-6
Condenser Calculated pHT using MULTEQ Amine Blend (with Acetate and Formate) and
Amine Blend (with CYC removed)

Sample Location Temp. Neutral ETA / NH3 NH3 pHT


deg C pHT 90% Steam 90% Steam Difference
Calculated pHT Calculated pHT
Condensate-1 Condenser 45 6.70 8.59 8.45 +0.14
Condensate-2 8.89 8.52 +0.37
Condensate-3 8.98 8.43 +0.55
Condensate-4 9.01 8.37 +0.64

The overall difference removing the CYC from the calculated results was only ~0.05 pH units
despite the CYC ppb concentration being higher than the ammonia and ETA. This illustrates the
importance of amine selection.
A similar analysis can be performed on the FFLP evaporator for this field study unit (Table 9-7
through 9-9).

Table 9-7
Steady State Chemical, Decomposition Product and Contaminant Levels

Sample ETA CYC Ammonia Acetate Formate Sodium Chloride Sulfate


(ppb) (ppb) (ppb) (ppb) (ppb) (ppb) (ppb) (ppb)
LP Drum-1 423 510 1650 61.5 17.2 17.2 1.3 1.52
LP Drum-2 765 934 2210 42 16.9 5.3 2.2 5
LP Drum-3 1030 1400 1260 44.5 23.2 15.4 1 0.6
LP Drum-4 1200 1470 1140 47.1 26 5.2 <0.5 1.3

9-18
Neutralizing and Filming Amine Treatment

As shown in Table 9-8, in each of the four cases superior at-temperature pH was achieved with
the neutralizing amine (even in the presence of the organic decomposition products) than with
ammonia alone (without the organic decomposition products) in the LP drum. Note while the
magnitude of the pH change was less than in the condensate, the impact of a pH increase at
150ºC (LP drum conditions) on iron solubility is much greater than at 50ºC (condenser
conditions). Note Table 9-8 gives the LP drum liquid phase pHT based on a calculated 90% water
10% steam, two phase condition.
Table 9-8
LP Drum Calculated pHT using MULTEQ Amine Blend (with Acetate and Formate) and
Ammonia Only (with no Acetate and Formate)

Sample Location Temp. Neutral ETA / CYC / NH3 NH3 pHT


deg C pHT 10% Steam 10% Steam Difference
Calculated pHT Calculated pHT
LP Drum-1 LP Drum 150 5.82 6.94 6.88 +0.06
LP Drum-2 7.03 6.94 +0.09
LP Drum-3 6.99 6.81 +0.18
LP Drum-4 6.99 6.79 +0.20

Table 9-9 re-examines the amine blend, this time removing the CYC to evaluate whether it was providing any
appreciable benefit to the at-temperature pH.
Table 9-9
LP Drum Calculated pHT using MULTEQ Amine Blend (with Acetate and Formate) and
Amine Blend (with CYC removed)

Sample Location Temp. Neutral ETA / NH3 90% NH3 pHT


deg C pHT Steam Calculated 90% Steam Difference
pHT Calculated pHT
LP Drum-1 LP Drum 150 5.82 6.91 6.88 +0.03
LP Drum-2 6.99 6.94 +0.05
LP Drum-3 6.93 6.81 +0.12
LP Drum-4 6.92 6.79 +0.13

The overall difference removing the CYC from the calculated results for the LP drums was again
only ~0.05 pH units despite the CYC ppb concentration being higher than the ammonia and ETA.
From this analysis it is clear that neutralizing amines even with the production of decomposition
products can result in improved at-temperature pH conditions and by extension the possibility of
improved iron corrosion control.

9-19
Neutralizing and Filming Amine Treatment

9.2.5.2 Decomposition Products Behavior in Field Studied HRSG


Using the data from the test unit it is possible to analyze the transport of amines and breakdown
products around the cycle. Figure 9-13 illustrates the flow path, temperatures and configuration
of the HRSG discussed at the opening of this subsection and analyzed above [3]. The test unit
had an equal blend of ETA and CYC injected downstream of the condensate discharge sample
point. Amine, ammonia and organic breakdown products were monitored at steady state
conditions at all the sample points illustrated. From this it is possible to estimate chemical
addition and breakdown rates.
               
         
         




 

 

   
     
      
  

 

           


          
       

Figure 9-13
Field Study HRSG: Flow path, Temperatures, Flows and Sample Points

As the condensate pump discharge sample (CPD) is prior to chemical injection, this sample point
represents the equilibrium amount of each chemical species surviving through the cycle. Since
the unit has an FFLP the total mass flow of chemical from the LP saturated steam and LP drum
water samples represents the total amount of chemical entering the cycle, based on the amount
that survived the cycle plus the amount of ETA and CYC added by the amine addition pump (see
Figure 9-13). Mass balances comparing ammonia and organic breakdown products at the CPD
plus the flue gas pre-heater return flow versus the LP saturated steam and LP drum water (after
accounting for the flue gas pre-heater recycle flow) should indicate no mass change if no ETA or
CYC are breaking down through this component. The additional amount of ETA and CYC in the
LP saturated steam and LP drum water versus the CPD should be equal to each other since the
amine blend addition is equal parts ETA and CYC. Table 9-10 examines this for the four
equilibrium tests examined. With the exception of test 3, each finds an approximately equal
addition of ETA and CYC on a mass balance basis (‘Difference’ rows in table), and essentially
no change in ammonia, acetate and formate mass flow rates. It can be assumed that test3 had
some level of analytical error resulting in the variance. For the other tests it is clear that ETA and
CYC are not breaking down from the condensate pump discharge point through to the FFLP
drum and LP saturated steam (~150ºC, 302ºF).

9-20
Neutralizing and Filming Amine Treatment

Table 9-10
Field Unit Mass Balance around Condensate and LP Drum

Flow per ETA CYC NH3 Acetate Formate


Sample Point
HRSG (lb/hr) (ppb) (ppb) (ppb) (ppb) (ppb)
CPD 580000 140 470 2470 51.1 16
From Flue Gas Pre-heater 42000 423 510 1650 61.5 17.2
1 FFLP Water 580000 423 510 1650 61.5 17.2
LP Saturated Steam 42000 0 3000 13600 0 1
Difference (lb/hr) 0.146 0.128 0.026 0.003 0.000
CPD 580000 380 1160 3410 33 12.2
From Flue Gas Pre-heater 42000 784 982 2210 42 16.9
2 FFLP Water 580000 784 982 2210 42 16.9
LP Saturated Steam 42000 11 8060 18500 0 0
Difference (lb/hr) 0.202 0.194 -0.012 0.003 0.002
CPD 580000 500 1900 2280 35.6 15.7
From Flue Gas Pre-heater 42000 1030 1400 1260 44.5 23.2
3 FFLP Water 580000 1030 1400 1260 44.5 23.2
LP Saturated Steam 42000 5.1 9500 10200 0 0
Difference (lb/hr) 0.308 0.109 -0.163 0.005 0.004
CPD 580000 740 1800 1740 46.1 26
From Flue Gas Pre-heater 42000 1200 1470 1140 47.1 26
4 FFLP Water 580000 1200 1470 1140 47.1 26
LP Saturated Steam 42000 54 9700 9350 1.2 1.9
Difference (lb/hr) 0.269 0.216 0.045 0.001 0.000

*Difference = Mass Flow from LP less Mass Flow from CPD

Table 9-11 examines mass balances around the IP and HP drums for the four tests. The mass
flow of amines, ammonia and breakdown products from the FFLP less the flow to the flue gas
pre-heater should equal the mass flow from the IP and HP saturated steam plus the IP drum
blowdown. The cascading blowdown from the HP drum to the IP drum is an internal flow and
does not figure in this mass balance (see Figure 9-13).
From Table 9-11 the ammonia (NH3) values are relatively unchanged with a slight increase in
each of the four tests. The CYC values are also relatively unchanged sometimes showing a slight
increase (which is not possible and is likely attributable to measurement errors). For ETA the
values are also relatively unchanged with only test 1 indicating any significant reduction. A
similar case exists for acetate. The one significant outlier is formate. In the case of formate there
is a significant decrease in the total formate exiting the IP/HP drum system from that entering it.
Across the four tests the average formate mass flow exiting the IP/HP drum system is 46% less
than the mass entering it. This indicates a significant level of formate decomposition in the IP/HP

9-21
Neutralizing and Filming Amine Treatment

drums, with ETA, CYC and acetate remaining fairly constant with little or no decomposition
occurring. At elevated temperatures (above 280ºC, 428ºF) formic acid is known to breakdown
preferentially to carbon monoxide and water via dehydration [20]:
HCOOH CO + H2O (Eq. 9-1)

This may account for the loss of formate through this section (in which case the decomposition
would be higher in the HP section than in the IP section due to the higher operating temperature).
Note this is only a possible mode of decomposition; it was not confirmed in the field work.
Table 9-11
Field Unit Mass Balance around IP and HP Drums

Flow per HRSG ETA CYC NH3 Acetate Formate


Sample Point
(lb/hr) (ppb) (ppb) (ppb) (ppb) (ppb)
FFLP Water 580000 423 510 1650 61.5 17.2
To Flue Gas Pre-heater 42000 423 510 1650 61.5 17.2
IP Saturated Steam 68000 450 420 1580 23.1 2.6
1
HP Saturated Steam 465000 280 440 1770 58.2 7.4
IP Blowdown 5000 2570 90 156 45.4 32.6
Percentage* 76% 85% 105% 87% 41%
FFLP Water 580000 784 982 2210 42 16.9
To Flue Gas Pre-heater 42000 784 982 2210 42 16.9
IP Saturated Steam 68000 880 940 2130 16 2.4
2
HP Saturated Steam 465000 690 1150 2400 42.7 12.9
IP Blowdown 5000 3080 93 220 40.5 33
Percentage* 94% 113% 106% 94% 70%
FFLP Water 580000 1030 1400 1260 44.5 23.2
To Flue Gas Pre-heater 42000 1030 1400 1260 44.5 23.2
IP Saturated Steam 68000 1280 1320 1220 18 3.2
3
HP Saturated Steam 465000 815 1560 1410 42 7.3
IP Blowdown 5000 5800 170 115 97 120
Percentage* 89% 108% 109% 89% 34%
FFLP Water 580000 1200 1470 1140 47.1 26
To Flue Gas Pre-heater 42000 1200 1470 1140 47.1 26
IP Saturated Steam 68000 1580 1460 1100 21.5 11.2
4
HP Saturated Steam 465000 1140 1580 1210 45 9.2
IP Blowdown 5000 7400 152 196 91.5 101
Percentage* 104% 106% 104% 90% 40%
Average Percent Change (of 4 Tests) 91% 103% 106% 90% 46%
*Percentage = Mass Flow FFLP less Mass Flow to Flue Gas Pre-heater over Saturated Steam plus Blowdown Mass
Flow

9-22
Neutralizing and Filming Amine Treatment

In the original study [3] an issue was noted with the hot reheat samples. These samples showed a
significant mass flow reduction in both ETA and CYC versus the IP/HP saturated steam, but also
in comparison to the CPD. The reduction compared to the IP/HP saturated steam can be
interpreted as breakdown through the IP/HP superheaters and the reheater. The reduction
compared to the mass flow at the CPD cannot be explained (even when accounting for the LP
saturated steam flow). As the CPD is before the chemical injection point there is no way for the
mass flow rate at the CPD to be greater than the mass flow from the reheat and LP saturated
steam flow. This discrepancy was interpreted as potentially indicating that the reheat sample
from the study was NOT representative of the actual reheat conditions due to thermal breakdown
of the amines within the sample line. This has important implications for steam monitoring when
applying neutralizing amines and may indicate that cooling at the point of sample extraction is
necessary to collect a representative high temperature reheat (or potentially superheat) steam
sample. In the field work, similar issues were seen on the other test units for reheat steam
samples [3]. Table 9-12 illustrates the discrepancy between the mass flow of ETA and CYC
between the CPD and reheat / LP saturated steam mass flow.
Table 9-12
Field Unit Mass Balance around Reheat / LP Saturated Steam and Condensate

Flow per HRSG ETA CYC NH3 Acetate Formate


 Sample Point
(lb/hr) (ppb) (ppb) (ppb) (ppb) (ppb)
      
      

      
Percentage* 189% 126% 93% 96% 52%
      
      

      
Percentage* 131% 103% 96% 82% 53%
      
      

      
Percentage* 160% 124% 108% 85% 49%
      
      

      
Percentage* 129% 112% 94% 99% 65%
Average Percent Change (of 4 Tests) 153% 116% 98% 90% 55%

*Percentage = Mass Flow Condensate over LP saturated Steam plus Reheat Steam Mass Flow

9-23
Neutralizing and Filming Amine Treatment

Since the reheat sample does not appear representative a mass balance of the IP/HP saturated
steam versus the CPD (less the LP saturated steam) can be used to give an assessment of the
survival of ETA and CYC across the steam section in the field study test unit. This presumes
neither are loss in significant quantities from the condenser air extraction system. Table 9-13
provides this analysis.
Table 9-13
Field Unit Mass Balance around IP/ HP Saturated Steam and Condensate (Less LP
Saturated Steam)

Flow per HRSG ETA CYC NH3 Acetate Formate


Sample Point
(lb/hr) (ppb) (ppb) (ppb) (ppb) (ppb)
LP Saturated Steam 42000 0 3000 13600 0 1
IP Saturated Steam 68000 450 420 1580 23.1 2.6
1 HP Saturated Steam 465000 280 440 1770 58.2 7.4
CPD 580000 140 470 2470 51.1 16
Percentage* 50% 63% 93% 104% 255%
LP Saturated Steam 42000 11 8060 18500 0 0
IP Saturated Steam 68000 880 940 2130 16 2.4
2 HP Saturated Steam 465000 690 1150 2400 42.7 12.9
CPD 580000 380 1160 3410 33 12.2
Percentage* 58% 56% 95% 91% 115%
LP Saturated Steam 42000 5.1 9500 10200 0 0
IP Saturated Steam 68000 1280 1320 1220 18 3.2
3 HP Saturated Steam 465000 815 1560 1410 42 7.3
CPD 580000 500 1900 2280 35.6 15.7
Percentage* 62% 86% 121% 99% 252%
LP Saturated Steam 42000 54 9700 9350 1.2 1.9
IP Saturated Steam 68000 1580 1460 1100 21.5 11.2
4 HP Saturated Steam 465000 1140 1580 1210 45 9.2
CPD 580000 740 1800 1740 46.1 26
Percentage* 67% 76% 97% 119% 298%
Average Percent Change (of 4 Tests) 59% 70% 101% 103% 230%

*Percentage = Mass Flow Condensate less LP saturated Steam over IP plus HP saturated Steam Mass Flow

From Table 9-13 on average 59% of the ETA survives through the IP/HP steam path to the
condenser. As there is virtually no ETA in the LP saturated steam flow this represents the
average equilibrium survival of ETA across the entire cycle for the four tests. For CYC, 70%
survives through the IP/HP steam path to the condenser. For CYC this is not the equivalent to the
survival across the entire cycle as a significant mass flow of CYC is present in the LP saturated
steam. Accounting for this the average equilibrium survival of CYC across the entire cycle for

9-24
Neutralizing and Filming Amine Treatment

the four tests is approximately 80%. Ammonia and acetate from Table 9-13 average around
100% survival, this indicates that while ammonia and acetate may be generated, at steady state
conditions these values are at equilibrium, meaning the generation of ammonia and acetate in the
cycle across the IP/HP steam path to the condenser, is in approximate balance with rejection of
both (presumably in the condenser). An alternate explanation would be that ETA and CYC are
not breaking down to acetate and ammonia in the steam path or that acetate if formed is further
breaking down, these remain open questions. For formate it is clear that there is a significant
increase across the IP/HP steam path to the condenser. The average formate mass flow in the
condensate is 230% of the IP/HP saturated steam flow. This suggests that one of the breakdown
products of ETA and CYC through this portion of the steam path is formate. While formate is
generated across this portion of the cycle as discussed previously a significant amount of formate
appears to decompose in the IP/HP drums. Recall from Table 9-11, on average this
decomposition was to 46% which is roughly the inverse of the generation across the IP/HP steam
path. This illustrates that formate across the entire cycle appears to have also reached an
equilibrium level as was the case for ammonia and acetate.
While formate is generated across the IP/HP saturated steam path to the condensate it is not
generated at a large enough rate to account for all the carbon loss from the ETA and CYC. To
examine the mass balance of carbon, an estimate of the carbon dioxide must be made. Degassed
cation conductivity was not measured during the field study, but with the anions measured it is
possible to estimate carbon dioxide levels from the measured straight cation conductivities, with
the presumption being that all the unaccounted cation conductivity is from CO2. Table 9-14 takes
the average values of the anions in the saturated steam samples and the condensate along with
the average cation conductivity and uses this to estimate the average CO2 concentrations.
Table 9-14
Field Study Unit Estimation of Average Carbon Dioxide in Saturated Steam and
Condensate (from Average Values across four tests measured)

Cation CO2
Acetate Formate Chloride Sulfate
Sample Point Conductivity Estimate
(ppb) (ppb) (ppb) (ppb)
(µS/cm) (ppb)
LP Saturated Steam 3.00 1.20 1.45 1.24 6.78 4300
IP Saturated Steam 0.64 19.65 4.85 1.30 1.73 205
HP Saturated Steam 0.87 46.98 9.20 0.95 0.66 270
CPD 1.40 41.45 17.48 1.85 3.13 700

With the estimated CO2 concentrations it is possible to do an average mass balance of carbon and
ammonia across the IP/HP steam path to the condensate. Table 9-15 illustrates the results.
It is clear that at equilibrium most of the carbon and ammonia generated through the continuous
decomposition of the ETA and CYC across the IP/HP steam path to the condensate is
unaccounted for (as would be expected at equilibrium). This suggests that these breakdown
products are in forms not measured, or that the formed ammonia and organic breakdown
products are being rejected at the same rate of formation (possibly in the condenser). Potentially
the unmeasured forms of breakdown products, if present are also well rejected in the condenser
(e.g. volatile organics with low water solubility), however these remain open questions.

9-25
Neutralizing and Filming Amine Treatment

Table 9-15
Field Study Unit Mass Balance IP/HP Steam Path to Condensate – Average Carbon and
Ammonia Values

Flow
Carbon
per ETA CYC NH3 Acetate Formate
Sample Point Dioxide
HRSG (ppb) (ppb) (ppb) (ppb) (ppb)
(ppb)
(lb/hr)
LP Saturated Steam 42000 17.5 7565 12912 0.3 0.7 4300
IP Saturated Steam 68000 1047.5 1035 1507 19.7 4.9 205
HP Saturated Steam 465000 731.3 1182 1697 47.0 9.2 270
CPD 580000 440.0 1332 2475 41.5 17.5 700
Percentage Change* 62% 73% 100% 104% 219% 162%
Measured
Measured Increase in Breakdown Un-
released from
Products (ppb) accounted
Amine (ppb)
Carbon (as C) 99.2 193.1 - 0.6 2.4 37.7 86%
Ammonia (as NH3) 70.3 45.6 2.1 - - - 98%

*Percentage = Mass Flow Condensate less LP saturated Steam over IP plus HP saturated Steam Mass Flow

This detailed analysis of decomposition across the field test unit demonstrates a number of
relevant points for this application of neutralizing amines:
 Acetate and formate reached equilibrium values in the cycle that are not strong functions of
the amine concentrations in the cycle or application rate.
– Formate formed across the steam path from the IP/HP saturated steam through the reheat
to the condensate. The HP superheat and the reheat section operated at 566ºC (1050ºF)
and respectively, 12.4 MPa (1800 psi) and 2.6 MPa (380 psi).
– Formate was lost (approximately 50% at equilibrium) through the IP/HP evaporator
sections. The HP evaporator operates at 327ºC (621ºF), 13.1 MPa (1900 psi), the IP
evaporator at 229ºC (444ºF), 2.9 MPa (420 psi).
– The mode(s) of removal of acetate and formate from the cycle that lead to each
equilibrium is unknown.
 Based on cation conductivity measurements and calculations, carbon dioxide was the most
abundant decomposition product retained in the cycle.
 Most of the carbon (average of 86% across the four tests) from decomposing ETA and CYC
was not accounted for in the measured acetate and formate concentrations and the estimated
carbon dioxide concentration (calculated from measured cation conductivity and anions).

9-26
Neutralizing and Filming Amine Treatment

 Most of the ammonia (average of 98% across the four tests) from decomposing ETA and
CYC was not accounted for in the measured ammonia concentrations
– Note there was no significant ammonia increase during the four tests, but all the ammonia
present in the cycle was from decomposition of ETA and CYC as no ammonia was
added. Essentially the measured concentrations represented equilibrium values based on
the ETA and CYC addition rate, decomposition rate and rejection of ammonia by the
steam / water cycle.
 The high temperature reheat steam sample, 566ºC (1050ºF), appeared to have significant
amine decomposition in the sample line.

9.3 Filming Amines Background Science


Filming amines applied in fossil and combined cycle plants are invariably proprietary chemicals
as such a general discussion of their properties is more difficult than in the case of neutralizing
amines which are for the most part generic chemicals. Filming amines differ from ammonia and
neutralizing amines in that they do not rely on reducing iron solubility through pH control.
Rather filming amines act by producing a physicochemical barrier on metal surfaces, essentially
by filming them. This barrier prevents water from coming into contact with the metal surface
thereby preventing the formation of corrosion cells.
Some key points for filming amines include:
 The filming amine needs to penetrate deposits and get to the metal surface.
 Filming amines that sit on the surface of a deposit may not provide much of a barrier.
 Filming amines take time to develop a protective film and also establish a residual in the bulk
solution.

9.3.1 Filming Amine Corrosion Protection Mechanism


Filming amines, often referred to as polyamines or fatty amines, are long chain hydrocarbons
belonging to the oligoalkyamino fatty amine family that have one hydrophobic end and one
hydrophilic end. Filming amines form a monomolecular “film” on the metal surface, creating a
physical barrier that prevents the water from reaching the steel surface and thus aids in the
protection of condensate/feedwater piping and steam generating equipment. The hydrophobic
alkyl group of the amine makes the metal surface unwettable and once formed, a protective film
remains intact even after the dosage has stopped [21]. Filming amines protect against oxygen and
carbon dioxide corrosion by replacing the loose oxide scale on metal surfaces with a very thin
amine film barrier.
The general chemical formula for filming amines is R1-[NH-R2]n-NH2, where n is an integer
between 0 and 7 and R1 is an unbranched alkyl chain with 12 to 218 carbon atoms and R2 is a
short chain akyl group that usually contains 1 to 4 carbon atoms. The two most common filming
amines are octadecylamine (ODA) (ODA, n = 0, R1 = C18H37) [15], and ethoxylated soya amine
(ESA).

9-27
Neutralizing and Filming Amine Treatment

9.3.1.1 Laboratory Results: Inhibition of Pitting and Crevice Corrosion


Inhibition of the corrosion of low pressure steam turbine steels, such as 1018 mild steel, 304, and
410 stainless steels, was investigated by EPRI in Reference [22] due to the importance of
corrosion in the thermal power industry. Three corrosion inhibitors, NALCO-2857, VpCI-357,
Anodamine, were evaluated to explore the effectiveness of the inhibitors for protecting turbine
blade and disc steels from corrosion and to evaluate the propensities of the inhibitors to induce
depassivation of the steel. Both the NALCO-2857 and Anodamine product are filming amines.
Inhibitor efficacy studies were carried out in closed vessels with the specimens being exposed to
an aqueous solution of the inhibitors to determine general corrosion rates by immersion test. The
distribution in pit depth (“damage function”) after immersion test was measured to characterize
the extent of damage by using confocal laser scanning microscopy. In parallel with these studies,
electrochemical studies of the inhibitors on the corrosion behavior of the steels were carried out.
From electrochemical impedance spectroscopy studies, the properties of the passive state, and
the impact of the inhibitor, were defined. Creviced specimen were fabricated so that the coupling
current between the crevice and the external surface could be measured to determine if a
properly applied protective barrier resulting from the application of film forming corrosion
inhibitors provides corrosion inhibition to creviced areas. The principal findings were as follows:
 The corrosion rate of 1018 mild steel decreased with more concentration of the inhibitors in
deionized (DI) water at 50°C (122°F) and 20 ppm Anodamine solution in DI water showed
improved protection against corrosion for 1018 mild steel. Likewise, Anodamine corrosion
inhibitor showed better protection against corrosion of 304 and 410 stainless steels.
 An improved corrosion protection was observed in 1018 mild steel when it was polarized in
Anodamine at + 20 mV (versus open circuit potential) for 87 minutes prior to the
electrochemical studies. A similar improvement in corrosion inhibition was observed in 1018
mild steel when a protective film was formed with other corrosion inhibitors prior to the
corrosion experiments.
 From the confocal laser scanning micrographs analysis, it was noted that the addition of
inhibitors helped in reducing the pit size and depth in both 1018 mild steel and 304 stainless
steel. 1018 mild steel immersed in DI water with 10 ppm NALCO-2857 showed relatively
small sized circular pits compared to that exposed in DI water without any corrosion inhibitor
at the same exposure time and temperature and the average pit depth decreased from 24 to 19
µm with the inhibitor addition. In addition, the number of pits was much lesser on 304
stainless steel exposed to inhibited DI water compared to that exposed to uninhibited DI
under similar conditions.
See Reference [22] (Inhibition of Pitting and Crevice Corrosion by Filming Amines and
Vapor Phase Corrosion Inhibitors. 1023065) for details on the test conditions for this study
and results.

9.3.1.2 Field Results: Hydrophobic Properties and Corrosion Product Transport


Figure 9-14 shows examples of a condenser hotwell and superheater tube filmed with a
proprietary amine (similar photos can be found in Reference [7] reviewing Russian experience
with film forming amines in HRSG plants).

9-28
Neutralizing and Filming Amine Treatment

Figure 9-14
Filmed Internal Surface Condenser Hotwell (Left), Superheater Tube (Right) [4]

Though results may vary depending on application and type of proprietary filming amine
applied, reports of improved overall iron and copper corrosion product transport have increased
in recent years [4-10]. Verib reported (see Figure 9-15) a significant decrease in iron corrosion
product transport after the application of a filming amine on a low-level continuous basis [4]
with similar results being reported by others [7-9].

350
Filming Amine Applied
300

HP Heater Drains
Total Iron ( g/kg)

250
LP Heater Drains

200 Economizer Inlet


Boiler Drum

150 Hot Reheat Steam

100

50

Figure 9-15
Iron Corrosion Product Monitoring Data, Drum Unit before and after Filming Amine
Application - Adapted from Reference [4]

9-29
Neutralizing and Filming Amine Treatment

9.4 Amine Potential Applications


If the use of an amine is being considered the following factors should be considered:
 There should be a reason; a specific corrosion problem that needs to be addressed that could
not be addressed with the optimization activities in Chapter 3.
 A significant benefit to the use of the amine should be able to be proven by testing; for
example a significant decrease in the amount of corrosion products generated in the cycle
and/or improved protection of equipment noted through visual observation or non-destruction
examination (NDE) testing.
 Monitoring should be implemented to routinely determine the amine concentration, amine
decomposition byproducts, corrosion products in the cycle and the influence on online
instrumentation.
 Other monitoring methods to detect contamination, must be employed because the
application of an amine will typically increase the cation conductivity, these may be limited
to degassed cation conductivity but could also include online ion chromatography.
 Point of chemical injection to maximize benefit.
 Safety, health and environmental approval to bring the chemical on site and treat / discharge.

9.4.1 Potential Use of Neutralizing Amines in Fossil and Combined Cycle Plants
Based on the research to date [1-3], general applications of neutralizing amines at the condensate
pump discharge will likely provide less benefit in conventional fossil plants due to the high
degree of decomposition through the power plant cycle, in particular the reheat section [1-3] than
could be achieved with an alternate dosing location. The net result is that relatively little of the
neutralizing amine is available for the intended use within the cycle, replaced by the thermal
decomposition products of ammonia, acetate, formate and carbon dioxide [1-3]. However this is
not necessarily the case in combined cycle plants with HRSGs, where there may be an
appreciable improvement in the two phase at-temperature pH conditions of the HRSG LP and IP
evaporator. In addition, while there is loss of neutralizing amine through decomposition in the
reheater, simulation and field results have nonetheless shown improvements in condenser
corrosion control results from the use of neutralizing amines (in some applications).
The following discussion identifies some of the areas that potentially could see an improvement
through the use of a neutralizing amine. Other drivers for the use of neutralizing amines not
discussed include:
 Customer restrictions on the use of ammonia,
 Copper alloy condensers,
 Saturated steam runs and returns to/from the customer,
 Systems with deaerators or process heaters with condensate return etc. (of particular concern
are designs where there is a phase change outside of the HRSG cycle).

9-30
Neutralizing and Filming Amine Treatment

9.4.1.1 Two Phase Flow-Accelerated Corrosion


Based on their distribution and dissociation properties many neutralizing amines should result in
a significant increase in the pH at operating temperature (pHT) under two phase flow conditions
such as those found in FFLP drums, SALP drums, IP drums and condensers. The application of
neutralizing amines at the condensate pump discharge is capable of achieving improvements in
these conditions, especially in the LP and IP drums. Thermal breakdown through the superheater
and reheater of combined cycle plant with HRSGs operating with steam temperatures at 538°C
(1000°F) or greater were found to range around 20-60% for ETA and CYC [3] which lessens the
effectiveness of amine addition at the condensate pump discharge for improving the pH at
operating temperature under two phase flow conditions in condensers. However, MULTEQ
simulations done for a field study unit which applied amines at the condensate pump discharge
found that the amines, and not ammonia and acetate from the thermal breakdown of the applied
amine, were the dominant factors in determining the two phase pH through the FFLP drum and
condenser (discussed in detail in Section 9.2.5) [3]. Note this was a particular test case result;
EPRI is currently working on additional research to understand the various mechanisms and
improve guidance on amine applications. Based on the current knowledge though, it is critical
for each plant to monitor corrosion control extensively when applying neutralizing amines to
understand if each individual unit is experiencing a benefit. In fossil plant applications from the
same study [3] the opposite was found to be the case when applying the neutralizing amine at the
condensate pump discharge. With the ammonia and acetate from thermal breakdown of the
applied amine the dominant factors in determining fossil plant condenser pH conditions under
two phase conditions. As such, while amine application at the condensate pump discharge may
result in enhanced corrosion control in two phase environments in a combined cycle plant with a
HRSG this must be tested and verified on a case by case basis through corrosion product
transport monitoring. Laboratory studies by Lister et al [23] at 200°C (392ºF) confirmed the
beneficial effect of ETA on the pH of water in two phase mixtures, with the benefit over
ammonia increasing with the steam voidage (volume percentage as steam).
For HRSG units with a FFLP drum which provides feedwater to the IP and HP evaporators the
application of a neutralizing amine has the potential to improve the pH of the feedwater over
what is achievable with ammonia alone due to the preferential loss of ammonia through the
FFLP drum (see Table 3-8). This improvement in pH should result in enhanced corrosion control
through the IP and HP economizers. However assessment of whether this improvement actually
occurs with the application of a neutralizing amine must be done on a case by case basis with
corrosion product monitoring.

9.4.1.2 Steamside Corrosion in Air Cooled Condensers (ACC)


Many neutralizing amines based on their distribution and dissociation properties could improve
two phase pH conditions within ACCs. This should reduce steamside corrosion in these
components, and field work reported by Stroman [24] has confirmed a reduction in corrosion
product generation after the application of a blend of ethanolamine (ETA) and ammonia, versus
operation with only the addition of ammonia (fed after the condensate polisher). In high pressure
units with superheat and reheat temperatures greater than or equal to 538°C (1000°F) this may
require that the neutralizing amine be applied in the LP Turbine or exhaust duct from the LP
Turbine to the ACC to prevent thermal decomposition from eliminating the neutralizing amine
prior to introduction into the ACC. A unit with a full flow condensate polisher applying a blend

9-31
Neutralizing and Filming Amine Treatment

of neutralizing amine and ammonia would be expected to have improved corrosion control
results versus a unit applying simply ammonia. As any anionic breakdown products that formed
could be readily removed. If the unit does not have a full flow condensate polisher the formation
of acetate and other neutralizing amine decomposition products may negate all or some of the
benefits from the application of the amine.

9.4.1.3 Steam Turbine Phase Transition Zone


The application of neutralizing amines directly to the IP-LP Turbine crossover may result in
improved early condensate and liquid film pH conditions. This improvement though might be
negated if the unit does not have a full flow condensate polisher by the presence of thermal
breakdown products of the amine such as acetate and formate. The application of an amine prior
to the superheater / reheater in units with steam temperatures greater than or equal to 538°C
(1000°F) may have a net neutral or potentially a net negative effect on the pH of early
condensate due to the formation of thermal breakdown products and relatively little of the amine
surviving the cycle to reach the phase transition zone of the turbine.

9.4.2 Potential Use of Filming Amines in Fossil and Combined Cycle Plants
The current EPRI research on filming amines is investigating whether their application can
accomplish the following:
 Decrease the occurrence of stress corrosion cracking and corrosion fatigue in steam turbines.
 Decrease the occurrence of corrosion fatigue failures in boilers and heat recovery steam
generators (HRSGs).
 Provide protection of all metal surfaces immediately upon application and during shutdown
for all outages of indeterminate length.
 Provide continual unit availability for restart.
 Provide long-term unattended layup protection.
As previously indicated this work is ongoing however many field trials have now been reported
with varying degrees of success in reducing corrosion product transport [4-11].
There are potentially two different applications of filming amines. First there is a continual feed
treatment, possibly in combination with ammonia or other neutralizing amines as a replacement
to standard treatments. The second is periodic application of filming amines, with the express
purpose to provide enhanced shutdown protection. EPRI is currently investigating both types of
applications.

9.5 Amine Applications: Instrumentation Considerations


Decomposition products from amines can have a significant impact on monitoring considerations
in the combined cycle / HRSG plant. The potential breakdown of amines to carbon dioxide and
organic acids will add cation conductivity to steam and condensate measurements making the
identification of condenser leaks and other serious contamination events more difficult. Plants
applying amines incapable of meeting the cation conductivity steam limits require additional
instrumentation to ensure that corrosion damage does not occur. The following discussion covers
these requirements. It has been shown conclusively through both laboratory and field testing that

9-32
Neutralizing and Filming Amine Treatment

all neutralizing amines discussed in this chapter will undergo some level of breakdown (often in
excess of 50%) in fossil and combined cycle plants [1-3,19]. For filming amines this may not
always be the case. Some filming amine field applications [10] have noted no impact on cation
conductivity from the chemicals use. However this is not universally the case and will depend on
the proprietary filming amine applied along with the dosage and time of addition. Often times
filming amines are combined with a neutralizing amine, in these cases the decomposition of the
neutralizing amine will occur.

9.5.1 Unable to Meet Steam Cation Conductivity < 0.2 µS/cm


Plants applying amines that are unable to meet the steam cation conductivity target of < 0.2
µS/cm should retrofit the plant with degassed cation conductivity measurements at all or some of
the core monitoring locations for cation conductivity given in Table 4-1. For plants that opt to
only put degassed cation conductivity on a portion of these samples, it is advisable to have the
ability to cycle each of the core sample points to this meter during upset conditions. Table 9-16 is
an extract from Table 4-1 indicating the locations where cation conductivity monitoring is core.
Table 9-16
EPRI’s Core Cation Conductivity Monitoring Points (Adapted from Table 4-1)
(Note Shared Analyzers are permissible see Section 4.1.1)

Cation
Sample Point
Conductivity
Main Steam and Reheat Steam (MS/RH) and LP, IP, and HP Superheat Steam
LP, IP and HP Saturated Steam
LP, IP and HP Drum Evaporator Blowdown*
IP and HP Economizer Inlets / Feedwater
Condensate Polisher Outlet (CPO)
Condensate Pump Discharge (CPD)

*Degassed Cation Conductivity Monitoring is not required on Blowdown Water

Should the plant be able to maintain degassed cation conductivity levels < 0.2 µS/cm no further
online instrumentation should be required, as the degassed cation conductivity monitoring should
be sensitive enough to detect significant chloride and / or sulfate steam purity excursions.

9.5.2 Incapable to Meet Steam Degassed Cation Conductivity < 0.2 µS/cm
For plants applying amines where the degassed cation conductivity cannot meet the steam target
value, additional online instrumentation is warranted. For these plants the degassed cation
conductivity is not sensitive enough to be a reliable detection method for steam chemistry
excursions. In these cases online Ion Chromatography is strongly recommended for measurement
of steam chloride and sulfate levels. Failure to include this instrumentation leaves the steam
purity for anionic contamination essentially unmonitored. Should a contamination event occur
(such as a condenser leak) this puts the steam turbine at significant risk of corrosive deposits
forming due to the inability of the cation conductivity measurement to detect the upset due to the
masking presence of decomposition products from the amine(s).

9-33
Neutralizing and Filming Amine Treatment

Plants adopting online Ion Chromatography should place this analyzer on a sequencer allowing
for monitoring of each of the steam samples from a single instrument. As discussed in Chapter 4
a U.S. based eastern utility has reported successful use of an online ion chromatograph for
monitoring chlorides and sulfates down to 0.5 ppb [25]. The system did require weekly
maintenance (changing inline filters and topping up of reagents).

9.5.3 Impact on Measurements


Neutralizing amines have been used extensively in the nuclear and fossil power generation
industry. They do not impact measurements using conventional water / steam monitoring
equipment [1-3]. I.e. conductivity, pH meters, etc. measure correctly when neutralizing amines
are applied. They also do not appreciably change cation resin performance for measurement of
cation conductivities [1]. I.e. cation conductivity measurement is still accurate when neutralizing
amines are applied.
For filming amines this is not always the case and the use of a filming amine can impact the pH,
conductivity, cation conductivity, sodium, ORP, dissolved oxygen measurements as well as
analysis with ion chromatography.
Work by Lendi and Wuhrman [26] showed that different proprietary filming amines will have
different impacts on measurements. The important point from this work is that when applying a
filming amine the plant must seek information from the vendor on how their chemical impacts
online instrumentation measurements (including chemistry, flow, pressure and temperature
instruments). In addition, the plant chemistry staff must diligently review measurements to
ensure that the filming amine is not resulting in erroneous results.
Table 9-17
Impact of Three different Filming Amines on Various Online Measurements

Measurement Filming Amine 1 Filming Amine 2 Filming Amine 3


Specific Conductivity No impact Coating on probe lowering readings (corrected
after stopping addition)
pH* Drift No impact No impact Caused drift when
amine was dosed
pH* Stability No impact No impact No impact
Sodium* Step Response No impact No impact No impact
Sodium* Calibration No impact Not Tested Not Tested
Oxygen** Response Time No impact No impact No impact
ORP Response Time Very Slow Very Slow Very Slow
Cation Conductivity Reduced resin performance, increasing cation conductivities (does not correct
after stopping addition, resin had to be replaced)

*Ion Specific Electrode **Clark type probe

9-34
Neutralizing and Filming Amine Treatment

Table 9-17 gives the results for three different filming amines from the Lendi and Wuhrman
work. In general chemistry parameters that rely on surface responses (e.g. conductivity, ion
specific electrodes, ORP) would be expected to be more susceptible to interference from filming
amines than measurements that rely on colorimetric techniques (e.g. phosphate, silica). In
addition measurements that rely on ion exchange may be significantly impacted depending on
how the filming amine acts on the ion exchange resin used (e.g. cation conductivity
measurements, online or offline ion chromatography can be impacted if the filming amine coats
the cation resin changing its exchange properties).

9.6 Amine Treatment: Conversion / Optimization / Evaluation

9.6.1 Choosing an Amine


A general note in selecting an amine is that neutralizing amines have been investigated by the
nuclear industry and found in most cases not to have a significant impact on condensate
polishing performance [1]. For filming amines, this has often been found not to be the case.
Several proprietary filming amines were found to have a significant detrimental effect on the
performance of the polisher resin [27]. Although due to the proprietary nature of these chemicals
it is not possible to say that this is a universal property of filming amines and users are
encouraged to work with their chemical vendor to evaluate specific proprietary filming amines
when applying them in a plant with condensate polishing, ensuring that the polisher is out of
service when applying the chemical if the impact is not known.

9.6.1.1 Neutralizing Amine: FFLP Drum


To address two phase FAC in an FFLP drum and the depletion of ammonia downstream of the
FFLP drum the addition of a neutralizing amine can provide benefit. Based on their properties
four of the examined amines will improve conditions in these applications through different
modes of action:
 Ethanolamine (ETA) which has significantly superior distribution at 150°C (302°F) than
ammonia can potentially improve the mitigation of two phase FAC in an FFLP drum. (Note
5AP would be expected to have similar results however it is not typically commercially
available).
 Dimethylamine (DMA) which has significantly superior dissociation at 150°C (302°F) than
ammonia can potentially improve the mitigation of two phase FAC in an FFLP drum.
 3-Methoxypropylamine (MPA) which has better dissociation and distribution at 150°C (302°F)
than ammonia can potentially improve the mitigation of two phase FAC in an FFLP drum.
Where each would be preferred is dependent on the steam flow from the FFLP. For units with
higher steam flows (greater than 25% of feedwater flow) ETA which will preferentially remain
in the water phase and provides slightly greater at-temperature pH benefit would be preferred.
Below these steam flows the enhanced dissociation of DMA will result in a slight improvement
in at-temperature pH and in turn lower iron solubility than the other two amines. MPA will fair
equally well in both cases. Regardless, the improvement in pH versus what is achievable with
ammonia alone will be relatively minor when ammonia controlled pH is in excess of 9.8 at 25°C
in the feedwater as the ammonia loss through the FFLP to the steam will typically be about 50%

9-35
Neutralizing and Filming Amine Treatment

of the feedwater concentration resulting in a drop of ~0.1-0.2 pH units. With an ammonia


feedwater pH of 9.8, the resulting FFLP pH would be ~9.6 which corresponds to relatively low
magnetite solubility according to Figure 9-1 of less than 5 ppb. If ammonia based pH cannot be
maintained above 9.4 then application of a neutralizing amine to the equivalent of 200 ppb TOC
(see Table 9-3), will result in appreciable improvement in the FFLP bulk water at-temperature
pH and result in a corresponding reduction in magnetite solubility.

9.6.1.2 Neutralizing Amine: SALP Drum


To address two phase FAC in a SALP drum typically a solid alkali treatment will be effective.
However for units deciding to apply an amine treatment, amines with superior distribution to
ammonia will result in the greatest benefit to at-temperature pH. Both ethanolamine (ETA) and
5-Aminopentanol (5AP) would be the preferred products.
Table 9-18 compares the expected bulk water pH from ammonia alone at a pH25 = 9.6 in a SALP
drum, and then as a blend with each of the neutralizing amines to the equivalent 200 ppb as
carbon concentration given in Table 9-3. The clear benefit of ethanolamine (ETA) and 5AP is
demonstrated in the significant increase in pH150 (pH at 150°C) and corresponding reduction in
magnetite solubility (based on the at-temperature pH of ammonia) as given by Figure 9-1 (after
converting the at-temperature pH to an equivalent ammonia concentration using Figure 9-2).
Table 9-18
Neutralizing Amines addition effect on SALP (150°C) pH and Expected Magnetite Solubility
based on At-Temperature pH (Figure 9-2 and 9-1)

Feedwater SALP at Equilibrium Magnetite


NH3 Amine Solubility
Chemical (ppb) pH25 (ppb) (ppb) pH25 pH150 (ppb)
NH3 2286 9.60 389 - 9.11 6.6 < 40
+ DMA 376 9.65 366 182 9.18 6.7 < 30
+ ETA 509 9.63 281 5270 9.62 7.1 <5
+ 5AP 344 9.62 261 8379 9.83 7.3 <5
+ MPA 371 9.62 354 725 9.23 6.8 < 20
+ Morph 363 9.61 376 674 9.14 6.7 < 30
+ CYC 276 9.62 390 52 9.13 6.7 < 30

9-36
Neutralizing and Filming Amine Treatment

9.6.1.3 Neutralizing Amine: Condensers (Both Water and Air Cooled Condensers)
To address two phase FAC and increased general corrosion due to two phase conditions in
condensers, application of an amine with superior distribution to ammonia at temperatures from
30-100°C (86-212ºF) will result in the greatest benefit to at-temperature pH. Both ethanolamine
(ETA) and 5-Aminopentanol (5AP) would be the preferred products. The improvement to at-
temperature pH of first formed condensate (denoted in Figure 9-10 by the green area) from the
application of either ETA or 5AP is very significant and as such the resultant improvement in iron
corrosion control should be expected to be significant. Table 9-19 examines the improvement in
early condensate (10% water by mass / 90% steam) pH with ammonia alone and with various
amine blends to the 200 ppb as carbon TOC level. ETA and 5AP are able to increase the at-
temperature pH greater than possible with ammonia alone even when added to a pH of 10.
Table 9-19
Neutralizing Amines addition effect on Condensate (50 C) pH and Expected Magnetite
Solubility based on At-Temperature pH (Figure 9-2 and 9-1)

Feedwater / Condenser Early Condensate – 90%


Steam
Magnetite
NH3 Amine Solubility
Chemical (ppb) pH25 (ppb) (ppb) pH25 pH50 (ppb)
NH3 11,812 10.00 986 - 9.38 8.7 < 10

NH3 2286 9.60 295 - 9.03 8.3 < 20


+ DMA 376 9.48 236 397 9.20 8.5 < 15
+ ETA 509 9.63 162 3826 9.52 8.8 <5
+ 5AP 344 9.62 172 6938 9.76 9.0 <5
+ MPA 371 9.62 207 1694 9.31 8.6 < 10
+ Morph 363 9.61 260 1533 9.09 8.4 < 20
+ CYC 276 9.62 271 314 9.09 8.4 < 20

Improved corrosion product transport has been the finding from actual plant applications in ACC
reported by Stroman and Hawkins [24,28]. In this work HRSG plants equipped with ACC
switched from an ammonia only application (to pH ~9.2) to an ammonia and ETA blend, 4:1
application (to pH ~9.2). Table 9-20 illustrates findings from this experience.
The dramatic improvement in iron corrosion product transport can be examined using MULTEQ
simulation. Figure 9-16 models the two conditions for at-temperature pH of the water phase. In
the ammonia alone treatment the expected iron solubility based on at-temperature pH through the
two phase zone experienced by a majority of the tubing / surface area of the ACC would have a
magnetite solubility in excess of 60 ppb. Whereas, for the ammonia / ETA blend the expected
magnetite solubility based on the at-temperature pH should be less than 20 ppb. This aligns well
with the observed results in the field application.

9-37
Neutralizing and Filming Amine Treatment

Table 9-20
Field Application: Ammonia and Ammonia / ETA Blend in ACC [28]

Ammonia Ammonia | ETA (4:1)


Parameter pH = 9.2 pH = 9.2

Condensate Pump
> 100 ppb 10-20 ppb
Discharge Iron (ppb)

Inspection Results:
Steam Turbine Outlet
Ducting, South Wall
Steam By-pass

Inspection Results:
Steam Turbine Outlet
Ducting, Condensate
drain

Significant improvement in condensate iron as well as in the appearance of the ACC ducting
oxides occurred after conversion to the ammonia / ETA blend.

 Magnetite Solubility


 < 20 ppb


pH @ T





 Magnetite Solubility

 > 60 ppb

     


Steam Fraction

    

Figure 9-16
MULTEQ Calculation of At-Temperature pH with Ammonia only and Ammonia | ETA 4:1
Blend (pH25 = 9.2 fully condensed sample, graphed at 60ºC)

9-38
Neutralizing and Filming Amine Treatment

9.6.1.4 Filming Amine


Due to the proprietary nature of filming amines it is not possible to provide direction on selection
of any particular chemical. Plants which cannot get satisfactory results with traditional
treatments or with neutralizing amines are the most likely to see some benefit from the
application of filming amines. This may be limited to heavily two-shifted units or rarely operated
units which experience significant offline corrosion. Selection of an appropriate filming amine
should be done in conjunction with chemical supplier and the plant chemist is encouraged to get
as much information as possible from the vendor prior to applying the chemical to minimize the
potential for adverse impacts upon application.

9.6.2 Selecting an Initial Concentration / Dosing Location

9.6.2.1 Neutralizing Amines: Initial Concentration / Dosing Location


Examinations by EPRI have shown that neutralizing amines breakdown in fossil and combined
cycle power plant conditions [2,19]. The breakdown products include organic acids, non-ionic
organics and carbon dioxide. Of these breakdown products non-ionic organics are unlikely to
cause any appreciable effects in the cycle. Carbon dioxide should cause no significant corrosion
issues during operation but will contribute to cation conductivity if degassed measurements are
not made and will lower operating pH by neutralizing ammonia / amines. Finally organic acids,
primarily acetate and formate, will add cation conductivity and potentially significantly lower the
pH of early condensate in steam [29,30].
Due to the formation of breakdown products, the ideal concentration for application is sufficient
to achieve the at-temperature pH improvement without overly contributing breakdown products
particularly organic acids. This will be unit specific as it is not possible to accurately predict the
breakdown rate of neutralizing amines in any particular cycle or the ability of the cycle to
remove these breakdown products.
Table 9-21 gives conservative recommended initial blends for the various neutralizing amines
examined.
Table 9-21
Recommended Initial Ammonia / Amine Blends (Requires Optimization)

Blend Ammonia : Amine


(by mass as 100% chemical)
NH3 | ETA 4:1
NH3 | 5AP 5:1
NH3 | DMA 5:1
NH3 | MPA 5:1
NH3 | Morph Not Recommended for use
NH3 | CYC Not Recommended for use

Blends including a second neutralizing amine may yield additional benefit over single
neutralizing amine and ammonia blends. For initial concentrations when a second neutralizing
amine is included, simply maintain the ammonia to neutralizing amine ratio given in Table 9-21.

9-39
Neutralizing and Filming Amine Treatment

For example a recommended initial blend for ammonia (NH3), ethanolamine (ETA) and 3-
Methoxypropylamine (MPA) would be to begin with 4 parts NH3, 1 part ETA and 0.8 parts
MPA.
Morpholine (Morph) and Cyclohexylamine are not recommended as they demonstrated little
benefit in terms of at-temperature pH control versus the results obtained by ammonia (NH3) in
the simulated conditions for condensers, LP evaporators and IP evaporators.
Neutralizing amines are typically applied at the condensate pump discharge using the addition
point typically used for ammonia feed with no adjustments in equipment necessary. In the
combined cycle / HRSG plant this may not be the optimal location for addition, especially if
improvements in steam turbine phase transition zone or early condensate pH are the goal. In
these cases it is expected superior results would be obtained by feeding into the IP/LP crossover
pipe (for larger steam turbines) as this will allow for the neutralizing amine(s) on their first pass
through the cycle, to reach these locations directly without first transiting the superheater /
reheater where chemical thermal breakdown will occur.

9.6.2.2 Filming Amines: Initial Concentration / Dosing Location


As these are proprietary chemicals the initial concentration for application should be determined
in conjunction with the chemical vendor. Due to the potential for unintended effects from the
application of filming amines it is highly recommended that the user take a very conservative
approach when first applying a new proprietary filming amine and to seek reference material
from the vendor for similar applications. Note however that is important to feed enough filming
amine to obtain the suggested residual by the vendor. The dosing location for filming amines is
typically the condensate pump discharge (where the ammonia feed typically takes place).

9.6.3 Required Monitoring


Normal monitoring per AVT (see Tables 4-1 and 4-2 as well as Chapter 5) should be performed
when operating with a blended amine treatment.

9.6.3.1 Cation Conductivity


As per Section 9.5 cation conductivity monitoring may require augmentation with either
degassed cation conductivity measurement (if normal operation cation conductivity exceeds 0.2
µS/cm) or online IC measurement (if normal operation degassed cation conductivity exceeds 0.2
µS/cm). Failure to do so will put the turbine at risk of undiagnosed chloride or sulfate chemistry
excursions well in excess of Action Level 3 values which can lead to significant corrosive
deposits within hours. Assuming contamination events with equal chloride and sulfate
concentrations Table 9-22 illustrates the risk. In the case of normal steam cation conductivity
> 0.2 µS/cm an exceedance of the chloride and sulfate Action Level 3 limits of 8 ppb can occur
without detection.

9-40
Neutralizing and Filming Amine Treatment

Table 9-22
Concentration to Cation Conductivity Relationship at 25ºC for various Contaminants in
Ultrapure Water

Concentration – ppb (µg/L)


Cation Conductivity Action Level 3 Limit is 8 ppb for Cl and SO4
(µS/cm)
Chloride (Cl) Sulfate (SO4)
0.20 9 9
0.40 19 19
1.00 50 50

For cases where carbon dioxide has caused the exceedance of steam cation conductivity limits
the effect can be substantially removed by measurement of degassed cation conductivity (see
Appendix A).
For cases where acetate or formate result in exceedance of the steam degassed cation
conductivity limit knowledge of the normal acetate and formate level can be used to characterize
the potential contamination of chloride and / or sulfate based on the degassed cation conductivity
measurement, however online IC is recommended. In the absence of online IC, equation 9-2 can
be used to estimate cation conductivity from chloride and sulfate:
CC Cl SO 4 DGCC Measured CC HCOO CC CH 3COO 0.055 S / cm (Eq. 9-2)

Where CCCl-SO4 is the estimated cation conductivity from chloride and sulfate, DGCCMeasured is the
measured degassed cation conductivity measured in S/cm, CCHCOO is the cation conductivity
from formate (from Figure 9-17) and CCCH3COO is the cation conductivity from acetate (from
Figure 9-18). Equation 9-3 can be used as an approximation of equation 9-2:
CC Cl SO 4 DGCC Measured 0.0083 HCOO( ppb) 0.0055 CH 3 COO ( ppb) (Eq. 9-3)

Where CCCl-SO4 is the estimated cation conductivity in µS/cm from chlorides and sulfates,
DGCCMeasured is the measured degassed cation conductivity measured in S/cm, HCOO is the
formate concentration in ppb and CH3COO is the acetate concentration in ppb.

9-41
Neutralizing and Filming Amine Treatment

1.65

1.45

1.25
Conductivity ( S/cm)
1.05

0.85

0.65

0.45

0.25

0.05
0 50 100 150 200
HCOO ppb

Figure 9-17
Formate (HCOO-) concentration versus Conductivity (Cation) [18]
1.65

1.45

1.25
Conductivity ( S/cm)

1.05

0.85

0.65

0.45

0.25

0.05
0 50 100 150 200
CH3COO ppb

Figure 9-18
Acetate (CH3COO-) concentration versus Conductivity (Cation) [18]

9-42
Neutralizing and Filming Amine Treatment

9.6.3.2 Specific Conductivity and pH Relationship Applying Neutralizing Amines


The specific conductivity and pH relationship used for ammonia given in Equation 9-4 is applicable for neutralizing
amines with an expected error < 0.05 pH units in power plant applications [31,32]:

pH 25 8 .55 log( Conductivi ty 25 ) 0 .032 (T 25 ) (Eq. 9-4)

Where Conductivity25 is the specific conductivity at 25°C in S/cm, and T is the actual
temperature of the sample in °C. This relationship is applicable only if the non-degassed cation
conductivity is < 0.5 µS/cm [18], above this value the relationship breaks down due to the
presence of carbon dioxide (see Figure 4-9).

9.6.3.3 Precautions for Monitoring Applying Filming Amines


Per Section 9.5 proprietary filming amines can have significant impacts on several different
measurements including:
 Conductivity (both specific and cation)
 Ion specific electrode measurements (pH, Sodium)
 Oxidation Reduction Potential (ORP)
 Ion Chromatography (Cl, SO4)
For any plant choosing to apply a proprietary filming amine they need to assess the impact of the
filming amine on these measurements and come up with unit specific plans on how to address
these issues. As illustrated in Section 9.5 the impact of different filming amines on these
measurements vary. One potential solution for filming amines impact on instrumentation is to
include a filter in the sample line with a large surface area for the filming amine to be consumed
filming prior reaching sensitive cation exchange columns or measurement probes. Whether such
a technique can be applied successfully would require field testing.

9.6.4 Optimization
Baseline monitoring of all parameters, in particular cycle iron corrosion product transport is
required prior to initiating optimization activities. Average normal full load values for all
parameters given in Table 4-2 (core and diagnostic parameters) should be established. As well as
startup values if the unit has frequent startups. These provide the baseline to evaluate changes of
chemistry control as a result of chemistry treatment program changes. If these fall outside normal
control limits, optimization per Section 3.3 of Chapter 3 should be attempted prior to applying a
blended Amine Treatment. Table 9-23 can be used for completing the baseline values for all
recommended parameters when attempting to establish or optimize an Amine Treatment.

9-43
Neutralizing and Filming Amine Treatment

Table 9-23
Recommended Baseline Evaluation Table (Complete with Average Values)

Degassed Cation Conductivity - µS/cm


Conductivity (Specific) - µS/cm

Cation Conductivity - µS/cm

Dissolved Oxygen – ppb

Acetate (CH3COO) – ppb

Formate (HCOO) – ppb


Ammonia (NH3) – ppb
Sulfate (SO4) – ppb
Chloride (Cl) – ppb
Sodium (Na) – ppb

Iron (Fe) – ppb*


pH – pH units

Amine – ppb
Sample Point

Main Steam (MS)


Reheat Steam (RH)
HP Superheated Steam
IP Superheated Steam
LP Superheated Steam
HP Drum Blowdown
IP Drum Blowdown

LP Drum Blowdown
HP Economizer Outlet
IP Economizer Outlet
LP Economizer Outlet
HP Economizer Inlet
IP Economizer Inlet
LP Economizer Inlet
Deaerator Outlet
Condensate

*Iron (Fe) monitoring requires appropriate sampling and analysis techniques to get representative results, the reader
should refer to Appendix F. NOTE: grab iron samples taken for analysis with a spectrophotometer will NOT result
in accurate results unless the sample is digested and a method with a detection limit of 1 ppb or less is used (see
Appendix G).
Ion chromatography analysis for chloride, sulfate, acetate and formate samples as well as for amines and ammonia
will typically be required. Effort to reconcile measured anions and the measured cation conductivity should be
taken (see Appendix A for relationships).

9-44
Neutralizing and Filming Amine Treatment

9.6.4.1 Neutralizing Amines


STEP 1: Switch to Amine / Ammonia Blend
Leave all control values and limits per optimized control limits established by Section 3.3
optimization activities. Replace ammonia only feed with selected initial blend from Table 9-21.
STEP 2: Monitor new Baseline
Over next several weeks establish new normal full load and / or startup values for unit for all
parameters given in Table 9-23 with the priority parameters being iron at all monitoring points
(see Appendix F for proper sampling and analysis) and degassed as well as straight cation
conductivity values at the main steam (MS) and reheat steam (RH). The average values of all
parameters given in Table 9-23 can then be used for advanced evaluation of the program.
STEP 3: Optimization
Table 9-24 can be used for optimization based primarily on iron corrosion product monitoring of
the feedwater and drum water as well as the main steam / reheat steam degassed cation
conductivity (note sodium monitoring may be key to identify more significant steam excursions).
Table 9-24
Amine Treatment Optimization Table

If iron And if MS/RH


corrosion Degassed And if the And the
Then: Because:
product Cation unit: next step is:
transport has: Conductivity
Decreased from < 0.2 µS/cm Increase Further iron STEP 1 with
baseline or Amine portion corrosion product stronger
stayed the of blend and transport may be amine blend
same re-evaluate achievable

> 0.2 µS/cm Has an online Increase Further iron STEP 1 with
IC for steam Amine portion corrosion product stronger
monitoring Cl of blend and transport may be amine blend
and SO4 re-evaluate achievable
Does not have Decrease There is an STEP 1 with
an online IC Amine portion unacceptable risk weaker
for steam of blend and of undiagnosed amine blend
monitoring Cl re-evaluate steam purity
and SO4 upsets*
Increased from Try different Blend applied Return to
baseline formulation resulted in former
aggravating iron treatment or
corrosion control try different
blend and
return to
STEP 1
*
For units that ignore this recommendation it is advisable to use equation 9-3 in conjunction with the baseline steam
degassed cation conductivity, formate and acetate values to determine a) that the degassed cation conductivity is
from formate and acetate, i.e. that equation 9-3 = ~0.055µS/cm, for these units establishment of an action level limit
0.1-0.2 µS/cm above the degassed cation conductivity value determined from the baseline acetate and formate
monitoring should be established and regular chloride and sulfate sampling of the steam conducted.

9-45
Neutralizing and Filming Amine Treatment

9.6.4.2 Filming Amines


The process for filming amines is nearly identical however additional precautions will need to be
taken which depending on the filming amine applied may include taking condensate polishers
out of service and establishing contingency plans for any measurements which have been found
to be adversely impacted by the filming amine in pre-application testing or based on vendor
information. Vendors may have test procedure for monitoring residuals. Typical process would
be to feed at a low level and attempt to maintain chemistry parameters within normal range.
Depending on the cleanliness of the system and the feed rate it will take time to achieve the
recommended residual at which the feed should be able to be decreased to a “maintenance” dose.

9.6.5 Evaluation
The primary evaluation of the effectiveness of an Amine Treatment is the iron corrosion product
transport monitoring. Achievement of the < 5 ppb iron in HRSG drums and < 2 ppb iron at
HRSG economizer inlets/outlets indicate fully optimized corrosion control. This achieved with
less than 0.2 µS/cm steam degassed cation conductivity would represent optimal conditions.
Further evaluation of the treatment should be performed using the criteria in Section 3.4 and per
evaluation of inspection results prior and post conversion per Appendix B for the HRSG, in
terms of oxide appearance and coverage. Table 9-20 showed an example of the oxide re-growth
where a neutralizing amine treatment had a beneficial effect in an ACC. Figure 9-19 gives the
same for a WCC using a blend of ammonia, ETA and MPA.

Figure 9-19
Wet Cooled Condenser: Oxide Re-growth with Optimized Ammonia / ETA / MPA Treatment
(Right Side) versus Un-Optimized Ammonia / ETA (Left Side) – [33]

9-46
Neutralizing and Filming Amine Treatment

In addition to evaluating corrosion product transport and inspection results HP evaporator tube
samples pre and post amine treatment conversion should be taken for comparison versus each
other as well as versus Figure 3-3 (chemical cleaning requirement based on deposit weight
density). Industry surveys have indicated that amine blend treatments have resulted in significant
increases in HP evaporator deposits weight density. However these surveys have only included
HRSGs where the amine blend included a reducing agent [34]. Limited tube results from units
applying an amine blend without a reducing agent have shown deposit weight densities well
below cleaning requirements (see Figure 9-20).

Before bead blast cleaning left side and after right side.

Figure 9-20
HRSG HP Evaporator Tube Analysis (12 years of Service) – Unit with ACC and Optimized
Ammonia / ETA Blend Treatment (no Reducing Agent) [35]

9-47
Neutralizing and Filming Amine Treatment

When a filming amine is applied there may be beading of water droplets noted upon inspection
when it is sprayed onto filmed surfaces as well as on tube samples (see Figure 9-14 for
examples). However the evidence of water beading may or may not be a good indication for the
film protection. This may depend on the oxide thickness as well and whether the surface has the
opportunity to dry out. As such, inspections looking for pitting damage, caused by offline
corrosion, are advised for a better indication of protection.

9.6.6 Control Limits


Control limits for a unit on an Amine Blend Treatment should be similar to those for AVT (see
Chapter 5), modified by the optimization activities in Section 9.6.4. These optimization activities
may lead to different pH / specific conductivity control limits based on iron corrosion control
results and steam cation conductivity.
As discussed above compliance with feedwater and steam cation conductivity limits may be
compromised in which case additional monitoring as described in this chapter are necessary to
ensure steam purity requirements are met. For plants which are using online IC the chloride and
sulfate steam target values are the primary control and not the steam cation conductivity.

9.7 References
1. Assessment of Amines for Fossil Plant Applications. EPRI, Palo Alto, CA: 2010. 1017475.
2. Thermal Degradation of Amines in Supercritical Water. EPRI, Palo Alto, CA: 2010.
1021499.
3. Interim Guidance - Amine Treatments in Fossil Power Plants. EPRI, Palo Alto, CA: 2010.
1019636.
4. G.J. Verib “An Alternative Chemistry for Both Operational and Layup Protection of High-
Pressure Steam=Water Cycles Using an Organic Filming Amine” PowerPlant Chemistry
2011, 13(5).
5. W. Hater, N. Rudschutzky, D. Olivet “The Chemistry and Properties of Organic Boiler Feed
Water Additives based on Film-Forming Amines, and their use in Steam Generators” EPRI
Major Component Reliability Workshop, Barcelona Spain, 2011.
6. A. Bursik “Polyamine/Amine Treatment – A Reasonable Alternative” PowerPlant Chemistry
2004, 6(9).
7. S. Yu. Suslov, A.V. Kirlina, I.A. Sergeev, E.A. Sokolova, and I.S. Suslov, “Experience with
Amines in Russia” Proceedings Third International Conference Interaction of Organics and
Organic Plant Cycle Treatment Chemicals with Water, Steam and Materials, Heidelberg,
Germany May 2012.
8. K.J. Galt and S. Sulliman, “Use of Filming Amine Treatment for Cycle Chemistry Operation
on an 8.4 MPa Drum Boiler Unit after 20 years of Storage” Proceedings Third International
Conference Interaction of Organics and Organic Plant Cycle Treatment Chemicals with
Water, Steam and Materials, Heidelberg, Germany May 2012.

9-48
Neutralizing and Filming Amine Treatment

9. G. Schreiber, D. Dreßler, A. de Bache, H. Choschzick, W. Hater “Changing Water-Steam-


Cycle Treatment from Oxygen-Binding / Ammonia to Film Forming Amines at Combined
Cycle Power Plant Jena” Proceedings International Conference Combined Cycles with
HRSGs – Heat Recovery in the Industry Holiday Inn Hotel, Heidelberg (Germany), May 6–8,
2013.
10.G. J. Verib, “Experiences with a Filming Amine Program in High Pressure Fossil Utility
Boilers” Proceedings Third International Conference Interaction of Organics and Organic
Plant Cycle Treatment Chemicals with Water, Steam and Materials, Heidelberg, Germany
May 2012.
11.M. Beck, K. Walker, B. Workman “American Electric Power Filming Amine Trials”
Proceedings Tenth International Conference on Cycle Chemistry in Fossil and Combined
Cycle Plants with Heat Recovery Steam Generators, Seattle, Washington, June 26-28, 2012.
12.Flow-Accelerated Corrosion in Power Plants, EPRI, Palo Alto, CA: 1998, TR-106611R1.
13.R.B. Dooley, V.K. Chexal, “Flow-Accelerated Corrosion,” CORROSION/99, paper 99347,
NACE International, Houston, TX (1999).
14.G.J. Bignold, K. Garbett, R. Garnsey, I.S. Woolsey, “Tackling Erosion-Corrosion in Nuclear
Steam Generating Plant,” Nuclear Engineering International, June, 1981, p.37.
15.P. Berge, J. Ducreaux, P. Saint-Paul, “Effects of Chemistry on Erosion-Corrosion of Steels in
Water and Wet Steam,” Proceedings of Water Chemistry in Nuclear Reactor Systems 2,
BNES, London, p.19.
16.M. Bouchacourt, “Impact of Water Chemistry on Corrosion-Erosion in One-Phase and Two-
Phase Flow,” Proceedings of water Chemistry in Nuclear Reactor Systems 5, BNES, London,
1989 p.135.
17.P. Sturla. “Oxidation and Deposition Phenomena in Forced Circulating Boilers and
Feedwater Treatment.” 5th National Feedwater Conference. Prague, 1973 (in French).
18.Cycle Chemistry Instrumentation Validation: Fundamental Relationships of Cycle Chemistry
Parameters. EPRI, Palo Alto, CA: 2010 1019641.
19.J. Lee, G. L. Foutch “Investigation and Findings on the Thermal Stability of Neutralizing
Amines” Proceedings Tenth International Conference on Cycle Chemistry in Fossil and
Combined Cycle Plants with Heat Recovery Steam Generators, Seattle, Washington, June
26-28, 2012.
20.Martin Schmeißer, “Decomposition of formic acid” Chemnitz University of Technology
August 31, 2011
21.Hatter, W., Rudschützky, N., and Olivet, D., “The Chemistry and Properties of Organic
Boiler Feedwater Additives Based on Film-Forming Amines and Their Use in Steam
Generators,” PowerPlant Chemistry, 2009, 11(2), 90.
22.Inhibition of Pitting and Crevice Corrosion by Filming Amines and Vapor Phase Corrosion
Inhibitors. EPRI, Palo Alto, CA: 2012. 1023065.

9-49
Neutralizing and Filming Amine Treatment

23. C. Lertsurasakda, P. Srisukvatananan, L. Liu, Derek Lister, J. Mathews “The Effects of


Amines on FAC in Steam-Water Systems” Fossil FAC International Conference,
Washington D.C., USA. 2013 March.
24. W. Stroman. “Chemistry change to address FAC-like corrosion in ACC tubes” Presentation
Air Cooled Condenser User Group, Las Vegas, NV: 2009.
25.Personal Communication from Mark Sindaco, PPL Corporation. April 18, 2011.
26.M. Lendi, P. Wuhrmann “Impact of Film Forming Amines on the Reliability of Online
Analytical Instruments” Proceedings Tenth International Conference on Cycle Chemistry in
Fossil and Combined Cycle Plants with Heat Recovery Steam Generators, Seattle,
Washington, June 26-28, 2012.
27.M. Hasan, E. Walker, A. Kabir, A. Apblett, G. Foutch “Impact of a Proprietary Filming
Amine on Condensate Polisher Resin” Proceedings Tenth International Conference on Cycle
Chemistry in Fossil and Combined Cycle Plants with Heat Recovery Steam Generators,
Seattle, Washington, June 26-28, 2012.
28.Personal Communication from Neil Hawkins, Capital Power. April 2013.
29.J. Stodola, R. Svoboda, "Investigation of Early Condensate Chemistry". International EPRI /
VGB Conference on Steam Chemistry - Interaction of Chemical Species with Water, Steam,
and Materials during Evaporation, Superheating and Condensation, Freiburg / DE, June
1999.
30.MULTEQ Version 4.0 Desktop Application, EPRI, Palo Alto, CA: 2006. 1014414.
31.W. Hater, A. de Bache, “Considerations on Conductivity and pH in Water-Steam Cycles
Using Organic Cycle Chemistry” Proceedings International Conference Combined Cycles
with HRSGs – Heat Recovery in the Industry Holiday Inn Hotel, Heidelberg (Germany), May
6–8, 2013.
32.M. Lendi, H. Wagner, P. Wuhrmann “pH calculation by differential conductivity
measurement in mixtures of alkalization agents” Proceedings International Conference
Combined Cycles with HRSGs – Heat Recovery in the Industry Holiday Inn Hotel,
Heidelberg (Germany), May 6–8, 2013.
33.N.A. Parker, A.G. Howell, “Neutralizing Amine Usage and Associated Changes in the
Appearance of Iron Oxides in the Condenser” Proceedings International Conference on
Flow-accelerated Corrosion in Fossil, Combined Cycle/HRSG and Renewable Energy
Plants, Arlington, VA. March 2013.
34.Heat Recovery Steam Generator (HRSG) Deposits: State of Knowledge Report. EPRI, Palo
Alto, CA: 2009. 1017629.
35.Personal Communication from William Stroman, Capital Power. April 2013.

9-50

You might also like