Vous êtes sur la page 1sur 7

This is an open access article published under an ACS AuthorChoice License, which permits

copying and redistribution of the article or any adaptations for non-commercial purposes.

Article

pubs.acs.org/IECR

Effects of Temperature and Pressure on the Thermolysis of


Morpholine, Ethanolamine, Cyclohexylamine, Dimethylamine, and
3‑Methoxypropylamine in Superheated Steam
David H. Moed,*,† Arne R. D. Verliefde,‡ and Luuk C. Rietveld†

Faculty of Civil Engineering and Geosciences, Delft University of Technology, Stevinweg 1, 2628 CN Delft, The Netherlands

Particle and Interfacial Technology Group, Ghent University, Coupure Links 653, 9000 Gent, Belgium
*
S Supporting Information

ABSTRACT: Alkalizing amines such as cyclohexylamine and dimethylamine have great potential for protecting steam−water
cycles against corrosion, but their thermal stability is limited and anionic decomposition products are a concern because of
increased corrosion risk. In this study, morpholine, ethanolamine, cyclohexylamine, dimethylamine, and 3-methoxypropylamine
were exposed to temperatures of 500, 530, and 560 °C at pressures of 9.5, 13.5, and 17.5 MPa to investigate the effects of
temperature and pressure on amine thermolysis kinetics. The surface/volume ratio of the reactor tube was 0.4 mm−1, close to the
value for superheater tubes in steam−water cycles. All amines thermolyzed by first-order kinetics, with the exception of
dimethylamine. The Arrhenius constants Ea, ln(A), and Va were obtained from the experimental data for all investigated amines.
The effect of pressure on the thermolysis kinetics was less pronounced than in previous studies and was different for each amine.
Dimethylamine did not degrade below 20% and 10% at 500 and 530 °C, respectively, despite the application of longer retention
times, suggesting the synthesis might occur. Limited practical data showed some promise for the applicability of the model to
steam−water cycles. More plant data are needed to fully validate the model. In all cases, thermolysis of the amines led to the
formation of between 150 and 600 ppb organic acid anions. In most cases, the concentrations increased linearly with increasing
degradation percentage. Acetate and formate were found to be major degradation products, with some propionate and traces of
glycolate. Cationic degradation products were ammonia and some amines, meaning that the complete thermolysis of an amine
does not necessarily lead to acidic conditions.

1. INTRODUCTION Some experimental data on the thermal degradation of


The demineralized water used in steam−water cycles is amines are available, but most such data focus only on
chemically conditioned to raise the pH of the water and conditions found in nuclear plants,9−12 with data applicable to
reduce failures due to corrosion. One of the applied chemical fossil-fired plants being more limited. Mori et al.13 tried to
treatment methods is all-volatile treatment (AVT), which uses simulate the thermolysis of six amines under fossil-fired plant
ammonia to raise the pH of both the boiler (feed) water and conditions at higher temperatures with a boiler−superheater
the steam or condensate. Although AVT is commonly applied test loop. They considered the effect of the superheater on
for steam−water cycles (SWCs), AVT requires high water thermolysis by varying the superheater exhaust temperature.
purity1 and still enhances the corrosion of any copper that They concluded that MOR breakdown in the boiler was
might be present in an SWC. Also, because ammonia is more undetectable and started only from 500 °C onward in the steam
volatile than water, it offers limited protection to corrosion in phase, whereas ETA started degrading already at steam exhaust
areas where flashing occurs or the first droplets of condensate temperatures of 450 °C. The retention time was not
are formed.2,3 Alkalizing amines such as morpholine (MOR) mentioned, making any calculation of kinetics impossible.
and ethanolamine (ETA) offer a less volatile alternative to The importance of the superheater in the applicability of
ammonia,4 making the SWC more resilient to corrosion.5−7 alkalizing amines in fossil-fired plants was also pointed out by
Because ammonia corrodes copper, amines can also be a good Bull14 in a report detailing the results from a fossil-fired plant
chemical water treatment solution for systems with mixed survey on the use of amines.
metallurgy. Amines have been applied in nuclear power plants In another study, MOR was dosed (9 ppm) to investigate the
for decades, and some guidelines are available.3 However, the effects of pressure, temperature, and retention time on amine
use of amines in fossil-fired plants is under debate, because of degradation kinetics in a flow reactor, mimicking a super-
the higher operating temperatures and pressures employed. heater.15 MOR degradation kinetics were found to be first-
Because amines are known to have limited thermal stability, order, increasing with temperature and decreasing with
they can decompose to form organic acid anions, lowering pH
and thereby increasing corrosion risks.2,8 Potential formation of Received: December 12, 2014
organic acid anions during amine thermolysis is currently Revised: February 13, 2015
preventing the general acceptance of amines for application in Accepted: February 18, 2015
steam−water cycles. Published: February 18, 2015

© 2015 American Chemical Society 2606 DOI: 10.1021/ie504849v


Ind. Eng. Chem. Res. 2015, 54, 2606−2612
Industrial & Engineering Chemistry Research Article

Figure 1. Schematic of the flow reactor used during the experiments (from left to right): argon, bottle containing the solution of interest, HPLC
pump, fluidized sand bed (FSB), metal tube in the FSB, cooling spiral, and back-pressure regulator (BPR).

pressure. The tubes used in the flow reactor, however, were Table 1. The Investigated Amines, Temperatures, Pressures,
only 1.5 mm in diameter, making the surface/volume (S/V) and Retention Timesa
ratio (4.65 mm−1) much higher than in full-scale fossil-fired
amineb T (°C) P (MPa) t (s)
plants, where the tube diameters are typically at least 10 times
larger. The metal oxide layer on the surface of the tubes MOR 500 9.5 0−5
appeared to also influence the thermolysis kinetics. In a ETA 530 13.5 5−10
previous publication, we provided a relation between the S/V DMA 560 17.5 10−20
ratio and thermolysis kinetics and made a distinction between CHA 20−30
homogeneous (in the bulk steam) and heterogeneous (on the MOPA
a
metal surface) thermolysis.16 Although heterogeneous thermol- T, temperature; P, pressure; t, retention time of the steam inside the
ysis accounted for 82−92% of the value of the degradation rate tube. bEach amine has been tested at every temperature, pressure, and
constant k for tubing with an S/V ratio of 4.65 mm−1, the retention time listed.
contribution was only 6−17% at an S/V ratio of 0.4 mm−1. This
led to the conclusion that an amine thermolysis study is best argon, and 0.1 ppm carbohydrazide was dosed to eliminate any
conducted with larger metal tubes, preferably with S/V ratios of remaining traces of oxygen. A more detailed description of the
0.4 mm−1 or less, to approximate superheater tube sizes. setup and procedure was already published by Moed et al.16
In this study, the effects of temperature and pressure on the 2.2. Analyses. Analysis of organic acid anions was
thermolysis of five amines that are being applied in steam− performed using a Metrohm (Schiedam, The Netherlands)
water cycles [MOR, ETA, dimethylamine (DMA), cyclohexyl- 881 ion chromatography system. A Metrohm A Supp 16 4.0/
amine (CHA), and 3-methoxypropylamine (MOPA)] were 250 anion column was operated at 67 °C with eluent
studied in a flow reactor equipped with a Hastelloy C-276 tube containing 7.5 mM Na2CO3 + 0.75 mM NaOH in ultrapure
with an S/V ratio of 0.4 mm−1. The data were used to construct water. The suppressor was regenerated with 50 mM H2SO4,
a predictive model to assess the applicability of amines in high- and the limit of detection was 1 ppb. For cation analysis, a
temperature and -pressure steam−water cycles. The organic Metrohm C5 cation column was used, with 3 mM HNO3 as the
acid anions and ammonia that are produced during amine eluent, for which the limit of detection for MOR, ETA, MOPA,
thermolysis are discussed, and an explanation is given as to why and DMA was 5 ppb and that for CHA was 50 ppb. The
DMA does not degrade according to first-order kinetics. detection limit for ammonia was 1 ppb.
2.3. Temperature and Pressure Relations. First-order
isobaric degradation kinetics as a function of absolute
2. MATERIALS AND METHODS
temperature were modeled according to the equations17
2.1. Setup and Conditions. A schematic of the
experimental setup is provided in Figure 1. A high-performance r = k(T )C (1)
liquid chromatography (HPLC) pump sends the solution of
k(T ) = A e−Ea / RT (2)
interest through Hastelloy C-276 tubing inside a fluidized sand
bath at a fixed flow rate. The tube inside the fluidized sand bed where r is the degradation rate, k(T) is the degradation rate
was a 12.5-mm Hastelloy C-276 tube with a volume of 6.2 mL constant as a function of temperature, C is the concentration of
and an internal diameter of 10 mm, thus having an S/V ratio of the degrading compound, A is the pre-exponential factor, Ea is
0.4 mm−1. the activation energy, T is the absolute temperature, and R is
The amine solutions of interest were tested under the the universal gas constant.
conditions listed in Table 1, and the experiments were executed The latter equation can be written in logarithmic form as
in duplicate. Retention times were varied by adjusting the Ea 1
HPLC pump flow rate to 0.75, 1.0, 1.5, or 2.5 mL/min, and the ln(k) = ln(A) −
RT (3)
flow rate was checked for accuracy before every run. Because
water and steam increase in volume at higher temperatures, the Therefore, in an Arrhenius plot of ln(k) versus 1/T, the slope
retention time in the tubing is a function of flow rate, of the regression line through the data points gives the
temperature, and pressure inside the tube. Therefore, applied experimental value of Ea/R, and the y intercept corresponds to
retention times are listed in ranges. Oxygen was removed with ln(A).
2607 DOI: 10.1021/ie504849v
Ind. Eng. Chem. Res. 2015, 54, 2606−2612
Industrial & Engineering Chemistry Research Article

In a previous study, Moed et al.15 showed the independency steam. The linear regression lines given for each pressure
of the activation energy (Ea) and activation volume (Va) on enabled the calculation of the activation energy Ea (by using the
both temperature and pressure. When the natural logarithm of slope) and the prefactor A (by using the y intercept). The
the degradation rate constant (k) is plotted as a function of averages of the resulting three Ea and ln(A) values were used in
pressure (at a constant temperature), the activation volume can the model. The same data were used to plot ln(k) against P for
be derived from the slope each temperature (see Figure 3), from which Va was calculated
Va
ln(k) = ln(k ref ) − (P − Pref )
RTref (4)
where kref is the degradation rate constant corresponding to the
reference temperature (Tref) and reference pressure (Pref). The
calculated values for Va should be independent of the chosen
values for Tref and Pref. Combining eq 3 and 4, the reaction rate
as a function of temperature and pressure can be written as
⎡ E V ⎤
r = k(T , P)C = exp⎢ln(A) − a − a (P − Pref )⎥C
⎣ RT RT ⎦
(5)
To determine the experimental values for k, an integrated
first-order rate law was fitted to the experimentally determined
Figure 3. Arrhenius plot of the relation between ln(k) and P for the
decrease in concentration of the amines upon degradation. First thermolysis of ETA in dry steam at 9.5, 13.5, and 17.5 MPa.
order was assumed because it was found to be the case for
amine degradation in most previous studies.9,10,12,15,16,18,19
using the slope of the linear regression lines at each
C = C0e−kt (6) temperature. The average for the three temperatures was the
The reaction rate constant k can be found by plotting C/C0 value of Va used in the proposed model.
against time and calculating the exponential regression function The Arrhenius plots for the other four amines can be found
y = e−kx. The threshold for a good fit was chosen at R2 = 0.95, in the Supporting Information. Figures S5, S8, and S11 show
and when the results that did not meet this requirement, the the pressure relation of ln(k) for MOR, MOPA, and CHA and
corresponding experiments were repeated. Subsequently, A, Ea, have almost horizontal regression lines. Figures S4, S7, and S10
and Va can be determined from eqs 3 and 4 as described above. (Supporting Information) show their temperature relation of
ln(k), where the regression lines are close together. MOR,
3. RESULTS AND DISCUSSION MOPA, and CHA therefore show no clear pressure depend-
ency. All values found for Ea, A, and Va are summarized in Table
3.1. Temperature and Pressure Relations. Degradation 2, including the errors. The activation volumes of MOR,
curves were constructed from the obtained data. The MOPA, and CHA have been highlighted in italics, because of
thermolysis of all amines led to first-order regression curves, their apparent lack of pressure dependency.
with the exception of DMA, which will be discussed in more It must be noted that a higher activation energy does not
detail later. First-order kinetics applied to degradation necessarily imply higher thermal stability of a compound, as
percentages over 90% as well, showing concentration thermal stability also depends on the prefactor A. The
independency in the range of 1−9 ppm. This range could be activation energy indicates the amount of energy that is
larger, but the data cannot validate such a claim. The values for
k resulting from the degradation curves were used to create Table 2. Summary of Ea, ln(A), and Va Values Found for
Arrhenius plots for the temperature relation. An example of Each Amine, Including Standard Deviations (SDs) and
such a plot is given in Figure 2, which shows the relation Relative Standard Deviations (RSDs)
between ln(k) and 1/T for the thermolysis of ETA in dry
amine constant mean SD RSD (%) units
ETA Ea 189 1 1 kJ mol−1
ln(A) 25.5 0.4 2 s−1
Va 369 59 16 cm3 mol−1
MOR Ea 295 15 5 kJ mol−1
ln(A) 40.2 2.2 5 s−1
Va 137 134 98 cm3 mol−1
MOPA Ea 254 16 6 kJ mol−1
ln(A) 35.6 2.4 7 s−1
Va 6 151 2343 cm3 mol−1
CHA Ea 409 12 3 kJ mol−1
ln(A) 56.9 1.7 3 s−1
Va 272 112 41 cm3 mol−1
DMA Ea 100 3 3 kJ mol−1
Figure 2. Arrhenius plot of the relation between ln(k) and 1/T for the ln(A) 13.6 0.4 3 s−1
thermolysis of ETA in dry steam at 500, 530, and 560 °C. Va 346 19 6 cm3 mol−1

2608 DOI: 10.1021/ie504849v


Ind. Eng. Chem. Res. 2015, 54, 2606−2612
Industrial & Engineering Chemistry Research Article

required to start the degradation reaction, but the prefactor A 500 °C, thermal degradation stopped when DMA was at
determines how quickly the amine degrades as a function of around 20%, as shown in Figure 5B. At 530 °C, the DMA
temperature. For example, the thermolysis of CHA in dry steam concentration did not drop below 10%. At 560 °C, the DMA
has a high activation energy at 500 °C, resulting in a higher concentration went down to less than 2% within a few seconds,
value for k than for MOR, making CHA more stable than MOR but the compound never fully degraded. In some cases, DMA
at that temperature. However, the high prefactor determines concentrations measured for longer retention times were even
that the k values of CHA and MOR are almost the same at 560 higher than those measured for shorter retention times, and a
°C and 13.5 MPa (0.074 and 0.072 s−1, respectively). clear increasing trend was observed. For instance, at 500 °C and
A previous study showed much lower MOR stability than 17.5 MPa, 15% DMA remained after 14 s in the reactor. After
observed here.15 That study was executed with a smaller 43 s under the same conditions, 23% DMA was left. This
stainless steel tube (S/V = 4.65 mm−1), resulting in a large increase cannot be explained through the error of detection of
contribution of catalytic wall effects to the thermolysis. This the ion chromatography system.
emphasizes the importance of the S/V ratio when studying Longer retention times do not allow degradation of DMA
thermal degradation under conditions resembling those in full- below 20% at 500 °C, suggesting that the DMA thermolysis
scale plants, as was shown in ref 16. In the previous study, a reaction is reversible. This can happen only if DMA is
much greater effect of pressure for MOR was found as well. synthesized from its breakdown products. Mochida et al.20
This suggests that metal surface catalysis enhances the pressure investigated the synthesis of DMA and methylamine (MET)
dependency effect. from ammonia and methanol over metal oxides at 300−450 °C,
Another report discussed the thermal stability of MOR, which could yield up to 50%. A patent by Ashina et al.21
CHA, ETA, and DMA in an experimental setup similar to that describes almost the same process, but claims that higher
used in this work but with an S/V ratio of 4.65 mm−1.18 MOR, selectivity for DMA can be obtained by exposing methanol and
CHA, and ETA showed lower thermal stability in the smaller ammonia to steam at 250−700 °C. During the experiments,
tubing. ETA, for instance, underwent 90% degradation in 12 s ammonia was a degradation product, and the presence of
at only 490° and 17.5 MPa. DMA was an exception, showing methanol was plausible. MET was also found as a degradation
higher stability in the small tubing. DMA thermolysis, however, product in all DMA experiments, and its concentration
did not always follow first-order kinetics, thus making modeling remained constant at higher retention times. Although this
difficult, as discussed in the next section. reasoning is speculative, the synthesis of DMA in a superheater,
To compare the experimentally obtained values of k with or perhaps in the condensing stages, is plausible.
those calculated using eq 5 and the results in Table 2, the As mentioned before, DMA was much more stable in a 1.6-
experimental k values were plotted against the calculated k mm tube at 490 °C18 than in a 12.5-mm tube at 500 °C. This
values. The coefficient of determination R2 was used as a indicates that the synthesis of DMA benefits from a high S/V
measure for the goodness of fit. As an example, the relation ratio. Having relatively more metal oxide surface area would
between the measured and calculated values of k for ETA is benefit the synthesis of DMA, thereby increasing the apparent
shown in Figure 4. The figures for all other amines are provided DMA stability as the S/V ratio increases.
3.3. Organic Acid Anion Formation. Potential formation
of organic acid anions during amine thermolysis is currently
preventing the general acceptance of amines for application in
steam−water cycles. In a previous amine thermolysis study,16
formate and acetate were found to be the major anionic
degradation products of MOR and ETA. In addition, glycolate
was measured as a degradation product of MOR and ETA. In
the current study, however, propionate was also found in
samples from MOR degradation. All organic acid data from
MOR thermolysis experiments conducted in the current study
are summarized in Figure 6. Formate was the major
degradation product, which does not correspond to the
findings in the previous study with a higher S/V ratio. This
difference indicates that the surface of the tubes has an affect
not only on the degradation kinetics, but also on what
Figure 4. Calculated values of k plotted against the experimentally degradation product is preferably formed. Formate, acetate, and
obtained values of k for all experiments conducted with ETA. propionate all increased with increasing degradation (despite
the increasing temperature), whereas glycolate was found at
in the Supporting Information. The coefficients of determi- only trace levels throughout all MOR thermolysis experiments.
nation of the modeled compared to the experimental data for The other organic acid plots for all of the investigated amines
ETA, MOR, MOPA, CHA, and DMA were found to be 0.99, can be found in the Supporting Information. It can be seen that
0.97, 0.97, 0.92, and 0.99, respectively. formate and acetate from ETA degradation were found at
3.2. Dimethylamine. DMA requires special attention similar concentrations (by mass) and their concentrations
among the studied amines because it was the only compound increased linearly to 290 ppb for 100% amine degradation. The
that did not degrade by first-order kinetics in all situations. The glycolate concentration after ETA degradation seemed to
previously discussed Arrhenius plots and model calculations for decrease with increasing ETA degradation, but also decreased
DMA were made with k values derived from the shortest with increasing temperature, which indicates the low thermal
retention times (Figure 5A), but when DMA was exposed to stability of glycolate. For MOPA, the highest concentrations
longer retention times, first-order kinetics no longer applied. At found for the degradation products were those of formate (up
2609 DOI: 10.1021/ie504849v
Ind. Eng. Chem. Res. 2015, 54, 2606−2612
Industrial & Engineering Chemistry Research Article

Figure 5. Degradation percentage vs retention time for the thermolysis of DMA at 500 °C for (A) shorter and (B) longer retention times. First-order
degradation does not apply to longer retention times.

meaning that a large proportion of the amines degraded to


neutral compounds.
In all cases, thermolysis of the amines led to the formation of
organic acid anions. In most cases, the concentrations increased
linearly with increasing degradation percentage. Organic acid
anions are a concern in power-plant regions with two-phase
flow, because organic acids are more soluble than ammonia.
This means that, when amines are used, there might be a point
at which the acidity of the degradation products has enough
effect on the water phase that the solubility of magnetite (the
protective oxide layer found under power-plant conditions)
increases. In practice, not all amines would be dosed at the
same concentration, as their dosages will depend on their
Figure 6. Acetate, formate, glycolate, and propionate formation vs basicities. An amine that requires a lower dose to maintain
degradation percentage during thermolysis of MOR at 500, 530, and sufficiently high pH in a steam−water cycle will also form fewer
560 °C and 9.5, 13.5, and 17.5 MPa. organic acid anions. It also has to be noted that amines and
organic acid anions in a power plant will cycle through the
to 242 ppb) and acetate (up to 120 ppb). The propionate system, so the thermal stability of organic acids in superheated
concentration increased slightly with increasing MOPA steam matters as well. Regardless, the protection an amine
degradation, but did not exceed 17 ppb. offers is a combination of its thermal stability and its tendency
CHA produced only a low amount of organic acids, with the to produce organic acid anions.
highest total organic acid anion concentration being less than 3.4. Nitrogen Balance. As an amine breaks down, the
200 ppb. The acetate concentrations from CHA were slightly cationic degradation products could still offer some protection.
higher than the formate concentrations, which were slightly Ammonia and other cationic degradation products were
higher than the propionate concentrations. All three organic measured in this study to determine the percentage of nitrogen
acid anions increased proportional to CHA degradation. An that can be found as cations providing alkalinity. This quantity
offset of 5 ppb formate and 10 ppb acetate can be seen for low is expressed as a percentage of the total organic nitrogen in the
degradation percentages. This is caused by organic carbon influent. For MOR, which degraded into ammonia and small
contamination (from, e.g., the air or the interior of the plastic traces of ETA (up to 229 ppb), the results for the cationic
pump), which is hard to prevent, even in full-scale plants. This degradation products are shown in Figure 7. Not all nitrogen
is also the reason that DMA thermolysis seemed to produce
small amounts of acetate, which was not the case.
DMA produced almost exclusively formate, which can be
attributed to its small molecule structure, making the formation
of other organic acid anions improbable, unless synthesis
occurred. Formate was also observed to be the only anionic
degradation product by Bull14 when using DMA as an ammonia
replacement for a secondary circuit.
Although the acetate and propionate concentrations did not
fluctuate, the formate concentrations went up and down for all
amines, even under similar conditions and degradation
percentages. This can be seen mostly for ETA, MOPA, and
DMA, where the formic acid concentrations for high
degradation percentages varued in the ranges 159−271, 51−
242, and 62−299 ppb, respectively. What causes the formate
concentrations to be so unpredictable remains unclear. Over Figure 7. Percentage of total nitrogen measured as cationic
the entire study, 1−15% (by mass) of the carbon released by degradation products (theoretical and measured) for MOR at 500,
degraded amines could be traced back as organic acid anions, 530, and 560 °C and 9.5, 13.5, and 17.5 MPa.

2610 DOI: 10.1021/ie504849v


Ind. Eng. Chem. Res. 2015, 54, 2606−2612
Industrial & Engineering Chemistry Research Article

could be traced back as ammonia or other cationic degradation A difference in model and plant data was found in the
products. This is probably due to the degradation of the amines formation of organic acids in plant A, where ETA was dosed.
and/or ammonia to inorganic nitrogen or to the loss of Almost no formate was measured at this location, which is
ammonia (because of its volatility) during grab sampling. A different from what was measured in plant B and the
blank run with only 0.1 ppm carbohydrazide revealed that 33 observations during the ETA degradation in this study. Plant
(±15) ppb ammonia came from the carbohydrazide. The B had a formate/acetate ratio close to what was experimentally
results were corrected for this contribution. obtained in this laboratory-scale study. In plant C, the organic
For other amines, a higher degree of scattering can be seen in acid concentrations were very low, but MOPA was dosed at
the nitrogen recovery plots, which are shown in the Supporting only 20 ppb in this case.
Information. In some cases, 100% of the nitrogen can be
accounted for, usually at lower temperatures or lower 4. CONCLUSIONS
degradation percentages. In all cases, the dominant cationic This study has led to the following conclusions:
degradation product was ammonia. For DMA, traces of MET (1) All amines thermolyzed by first-order kinetics, with the
(up to 236 ppb) were found as a decomposition products as exception of dimethylamine.
well, whereas MOPA also degraded to some hydroxyl (2) The obtained Arrhenius constants are summarized in
propylamine (up to 204 ppb). Table 3.
With most nitrogen always being converted into ammonia or
another amine, the thermolysis of an amine in a power plant Table 3. Summary of Obtained Arrhenius Constants
does not necessarily have to lead to a hazardous pH drop, as the
pH effect of the formation of organic acids will mainly be Ea (kJ mol−1) ln(A) (s−1) Va (cm3 mol−1)
countered by the alkalizing effect of the ammonia. As stated in ETA 189 25.5 369
the previous section, this could still imply a lower degree of MOR 295 40.2 137
protection in two-phase flow, as some organic acids are less MOPA 254 35.6 6
volatile than ammonia and prefer the liquid film that covers the CHA 409 56.9 272
metal surface. DMA 100 13.6 346
3.5. Comparison with Practical Data. Practical data in
the literature that can be used to verify the model are almost
nonexistent, mostly because of the lack of information on (3) The effect of pressure on the thermolysis kinetics was
retention times in superheaters and reheaters. Some general much less pronounced than in previous studies, which can
comparisons can be made, however. Layton and Daniels22 probably be attributed to the low S/V ratio in this study, which
reported a power-plant survey on amine usage that included was applied to come closer to superheater conditions.
detailed data for ETA degradation in several parts of two plants (4) Dimethylamine did not degrade below 20% and 10% of
(A and B), and in plant B, CHA was also dosed. The third plant the initial concentration at 500 and 530 °C, respectively,
(C) ran (partially) on MOPA dosing. The superheat (SH) and despite the application of longer retention times. This suggests
reheat (RH) temperatures of plants A−C were 540, 566, and that synthesis might occur, either at high temperature or in the
538 °C, respectively, with pressures of 17.9, 13.1, and 13.1 condensing stages. Further studies are needed to clarify this
MPa, respectively. phenomenon.
Retention times could not be calculated from the reported (5) In all cases, thermolysis of the amines led to the
data, so degradation percentages can be compared only relative formation of organic acid anions. In most cases, the
to each other. Similarities were found between their study and concentrations increased linearly with increasing degradation
this study in the stability of CHA relative to ETA. Layton and percentage. Acetate and formate were found to be the major
Daniels22 showed a loss of CHA and ETA over the SH and RH, degradation products, with some propionate and traces of
but also showed that CHA is more stable than ETA at 551 °C. glycolate.
The average loss percentages of CHA and ETA were (47 ± (6) Cationic degradation products consisted of mostly
7)% and (60 ± 11)%, respectively, at a 551 °C RH steam ammonia and some amines, meaning that the complete
temperature. According to the model presented in this article, a thermolysis of an amine does not necessarily lead to acidic
47% decrease in CHA at 551 °C and a pressure of 13.1 MPa conditions. Not all nitrogen could be traced back to these
corresponds to a 70% loss of ETA under similar conditions. As cationic degradation products, suggesting that some other
such, the findings corroborate the practical data (within the inorganic nitrogen species were formed or that ammonia was
degree of accuracy of the plant data). When applying the same lost during sampling.
retention time to the conditions found in plant C (a heat (7) Comparison of the model to some limited practical
recovery steam generator similar to plant B), the model data measurements of amines in power plants shows some promise
predict a 70% loss of MOPA, which is close to the highest in the applicability of the model. A larger data set and more
degradation percentage of MOPA found in that specific plant of information on superheater retention times are necessary to
66%. validate the model proposed in this article, which should try to
It should be kept in mind that comparisons between model focus on retention times in the hottest sections of power plants.
data based on laboratory-scale experiments and full-scale data
need to be handled with care. Amines in a plant do not stay at a
fixed temperature for a few seconds, but follow a profile of

*
ASSOCIATED CONTENT
S Supporting Information
temperatures. Determining retention times and temperature Arrhenius, organic acid anion, and nitrogen recovery plots are
profiles in superheaters and reheaters is a challenge. However, provided to support the findings described in this article. This
making these kinds of comparisons is important for assessing material is available free of charge via the Internet at http://
the applicability of amine thermal stability modeling. pubs.acs.org.
2611 DOI: 10.1021/ie504849v
Ind. Eng. Chem. Res. 2015, 54, 2606−2612
Industrial & Engineering Chemistry Research Article

■ AUTHOR INFORMATION
Corresponding Author
International Conference on the Properties of Water and Steam, London,
U.K., Sep 1−5, 2013.
(19) Moed, D. H.; Verliefde, A. R. D.; Rietveld, L. C.; Heijman, S. G.
*E-mail: D.H.Moed@tudelft.nl. Tel.: +31-0-152786588. J. Organic acid formation in steam-water cycles: Influence of
Notes temperature, retention time, heating rate and O2. Appl. Therm. Eng.
The authors declare no competing financial interest. 2014, 65, 194−200.


(20) Mochida, I.; Yasutake, A.; Fujitsu, H.; Takeshita, K. Selective
ACKNOWLEDGMENTS synthesis of dimethylamine (DMA) from methanol and ammonia over
zeolites. J. Catal. 1983, 82, 313−321.
The authors thank EPRI and Evides Industriewater for financial (21) Ashina, Y.; Fujita, T.; Fukatsu, M.; Yagi, J. Process for producing
and technical support. Specific gratitude goes to Michael dimethylamine in preference to mono- and trimethylamines by gas
Caravaggio and James Mathews from EPRI for useful input and phase catalytic reaction of ammonia with methanol. U.S. Patent
discussions. 4,582,936, 1986.


(22) Layton, K. F.; Daniels, D. G. Interim GuidanceAmine
REFERENCES Treatments in Fossil Power Plants; EPRI: Palo Alto, CA, 2010.
(1) Dooley, R. B.; Shields, K. J.; Aschoff, A. F.; Ball, M.; Bursik, A.
Cycle Chemistry Guidelines for Operators and Chemists: All-Volatile
Treatment, Revision 1; EPRI: Palo Alto, CA, 2002.
(2) Svoboda, R. The Effect of Carbon Dioxide and Organics in a
Steam Turbine. PowerPlant Chem. 2009, 11, 20−29.
(3) Miller, A. D. PWR Advanced Amine Application Guidelines,
Revision 2; EPRI: Palo Alto, CA, 1997.
(4) Cobble, J. W.; Turner, P. J. PWR Advanced All-Volatile Treatment
Additives, By-Products, and Boric Acid; EPRI: Palo Alto, CA, 1992.
(5) Robinson, J. O.; Kluck, R. W.; Rossi, A. M.; Carvalho, L. Organic
chemical treatment of high-purity boiler feedwaterAdvantages and
limitations. PowerPlant Chem. 2009, 11, 690−699.
(6) Kluck, R.; Torres, J.; Antompietri, A.; Rivera, J. Experiences using
neutralizing amines to control pH and minimize FAC in combined
cycle plant. In Proceedings of the 71st Annual International Water
Conference; Engineers' Society of Western Pennsylvania (ESWP):
Pittsburgh, PA, 2010; pp 742−752.
(7) Lertsurasakda, C.; Srisukvatananan, P.; Liu, L.; Lister, D.;
Mathews, J. A. The effects of amine on flow-accelerated corrosion in
steam-water systems. PowerPlant Chem. 2013, 15, 181−189.
(8) Bellows, J. C. Hazards of Amine Use in Boiler Systems. Presented
at the 65th Annual International Water Conference, Pittsburgh, PA, Oct
17−21, 2004.
(9) Domae, M.; Fujiwara, K. Thermal decomposition of 3-
methoxypropy lamine as an alternative amine in PWR secondary
systems. J. Nucl. Sci. Technol. 2009, 46, 210−215.
(10) Shenburger, D. M.; Zupanovitch, J. D.; Walker, J. L. Loop
Testing of Alternate Amines for All-Volatile Treatment Control in PWRs;
EPRI: Palo Alto, CA, 1992.
(11) Feron, D.; Lambert, I. Thermal Stability of Three Amines in
Pressurized Water Reactor Secondary Systems. Laboratory and Loop
Experiments. J. Solution Chem. 1992, 21, 919−932.
(12) Gilbert, R.; Lamarre, C. Thermal Stability of Morpholine
Additive in the Steam-Water Cycle of CANDU-PHW Nuclear Power
Plants. Can. J. Chem. Eng. 1989, 67, 646−651.
(13) Mori, S.; Sato, T.; Shimura, Y., Characterization of Amines
under High Temperature Conditions and Their Use for Boiler Water
Treatment. Presented at Corrosion 2012, Salt Lake City, UT, Mar 11−
15, 2012.
(14) Bull, A., Dimethylamine as a Replacement for Ammonia Dosing
in the Secondary Circuit of an Advanced Gas-Cooled Reactor (AGR)
Power Station. Presented at the 16th International Conference on the
Properties of Water and Steam, London, U.K., Sep 1−5, 2013.
(15) Moed, D. H.; Verliefde, A. R. D.; Heijman, S. G. J.; Rietveld, L.
C. Thermolysis of Morpholine in Water and Superheated Steam. Ind.
Eng. Chem. Res. 2014, 53, 8012−8017.
(16) Moed, D. H.; Verliefde, A. R. D.; Rietveld, L. C. Role of Metal
Surface Catalysis in the Thermolysis of Morpholine and Ethanolamine
under Superheater Conditions. Ind. Eng. Chem. Res. 2014, 53, 19392−
19397.
(17) Laidler, K. J. Chemical Kinetics, 3rd ed.; Harper Collins: New
York, 1987.
(18) Moed, D. H.; Verliefde, A. R. D.; Heijman, S. G. J.; Rietveld, L.
C., Degradation kinetics of six alkalizing amines. Presented at the 16th

2612 DOI: 10.1021/ie504849v


Ind. Eng. Chem. Res. 2015, 54, 2606−2612

Vous aimerez peut-être aussi