Abstract
The corrosion behavior of 15CrMo used for water-wall tubes was studied in various urea-containing solution to determine the corrosion problem of water-wall tubes caused by urea in a coal-fired power plant. Urea decomposition tests, together with corrosion experiments, were carried out. The temperature was 320 °C, and the concentrations of urea were 70, 140, 280, 560 and 840 mg/L. Weight loss experiments and surface analysis indicated that the corrosion of 15CrMo steel is mainly manifested as localized corrosion. The corrosion rate of 15CrMo steel increased with the increase of urea concentration, and the maximum value reached 0.686 mm/y (mm per year) when the urea concentration was 840 mg/L. Electrochemical analysis suggested that the corrosion rates of 15CrMo were enhanced substantially by urea decomposition products. The results of UPLC-ESI-MS, infrared spectroscopy and X-ray diffraction showed that urea solution produced corrosive ions NH2COO− during the decomposition process, which caused the corrosion of 15CrMo. Results provided evidence as relevant explanation for the corrosion behavior of 15CrMo in urea solution.
1 Introduction
Since 2012, the regulations in China require that the emissions of coal-fired power stations contain less than 100 mg/m3 of nitrogen oxides, and that the emissions of natural gas-fired power stations contain less than 50 mg/m3 nitrogen oxides, which are the standard international levels. Denitrification has become essential for coal-fired thermal power plants, and selective catalytic reduction (SCR) is a widely used denitrification process (Bernhard et al. 2012; Kröcher et al. 2005). Ammonia (NH3) is usually used as the reducing agent in the SCR and reduces nitrogen oxides (NOx) into nitrogen (N2) and water (H2O) under the condition of catalyst to achieve denitration (Todorova et al. 2010). Urea, which is the NH3 storage compound for the SCR (Sahu et al. 2010), has a wide range of applications in the denitrification system for coal-fired thermal power plants because it is economical and safe (Eichelbaum et al. 2010a,b; Koebel et al. 2000). In SCR denitrification, the urea solution hydrolyses to produce NH3 and carbon dioxide (CO2) under a certain temperature and pressure, as shown in Equations (1) and (2) (Liu et al. 2014; Zhang et al. 2017). The main reaction processes where NOx is reduced by NH3 in flue gas are shown in Equations (3) and (4) (Eichelbaum et al. 2010a,b; Koebel et al. 2000):
In the above two reaction equations, the ammonium carbamate produced by the first reaction is corrosive to carbon steel. When the temperature exceeds 190 °C, urea has a corrosive effect on general stainless steel (such as 304 and 904 L) (Nockert and Norell 2014). The form of corrosion is mainly general corrosion, and large amounts of corrosion products are deposited (Shaikh et al. 2003). Nickel-based alloy (such as Inconel 718) is mainly dissolved and corroded with extremely low corrosion rate; titanium alloys form a continuous oxide film on the surface when the temperature exceeds 220 °C (Forzatti 2001; Lu et al. 2017).
15CrMo is a pearlitic heat-resistant steel that is used for water-wall tubes in coal-fired thermal power plants because of its high toughness, creep resistance and resistance to high-temperature oxidation and corrosion. A steam drum boiler of a coal-fired power plant was shut down because of failure of the water-wall tube (Figure 1). Before the failure, the water quality was determined to be abnormal, and a high concentration of ammonia-containing pollutants entered the system. After on-site chemical analysis and system troubleshooting, the pollutants were found to be derived from the urea solution used by the denitration system. During the hydrolysis of urea, some acidic substances will be generated. The hydrolysis products cause the original corrosion products to spall and migrate, thereby forming heavy deposits in areas where the flow is blocked. Stress corrosion cracking originating from intrinsic defects in the water-wall tubes under these heavy corrosion product deposits is the direct cause of the accident. Similar cases of corrosion caused by urea solution can be found in the literature studies (Lu et al. 2018; Starostin et al. 2016). Despite the industrial prevalence of such issues, few current basic research efforts are exerted in this area.

Cracks in water-wall tubes of coal-fired power plant.
In this study, the operational conditions of a coal-fired power plant were simulated in the laboratory, and the hydrolysis experiments of the urea solutions at different concentrations were carried out in an effort to better understand this corrosion phenomenon. Total organic carbon (TOC) analyses, ultraperformance liquid chromatography-electrospray ionization-mass spectrometry (UPLC-ESI-MS) and infrared spectra of the vapor and liquid samples were used to characterize the urea decomposition process. Weight loss measurements were used to establish the corrosion rate 15CrMo steel exposed to the decomposition products of the urea denitration solution. The impact of urea hydrolysis products on the corrosion behavior of water-wall tubes were studied employing electrochemical methods and microscopic characterization techniques which included metallographic microscopy, scanning electron microscope (SEM), energy dispersive spectrometer (EDS) and X-ray diffraction (XRD). Obtained results were applied to evaluate the corrosion mechanism of 15CrMo steel by urea hydrolysis products.
2 Materials and methods
2.1 Materials and solution
The urea used in this study was supplied by Datang Anyang Power Plant and contained a total nitrogen content of ≥46.4%. All of the simulation solutions were prepared using 18.25 MΩ cm ultrapure water. The 15CrMo coupons utilized in this study were 40 × 13 × 2 mm in size, and the alloy compositions in the components studied are given in Table 1.
Composition of the alloys in the studied components (wt %).
Alloy | C | Mn | Si | Cr | Mo | Ni | S | P |
---|---|---|---|---|---|---|---|---|
15CrMo | 0.12 ∼ 0.18 | 0.40 ∼ 0.70 | 0.17 ∼ 0.37 | 0.80 ∼ 1.10 | 0.40 ∼ 0.55 | ≤0.30 | ≤0.035 | ≤0.035 |
2.2 Urea solution decomposition experiment
The urea hydrolysis experiment was carried out with 600 mL of urea solution in an autoclave (F1.4-20/400, Kemao Experimental Equipment Co., Ltd, Dalian, China), simulating the operating condition in the thermal power plants. The experimental device was shown in Figure 2. The pH value of the solution in the pipe where the failure occurred (described in the introduction above) was between 10.0 and 11.0. At a pH of 10.6, the corresponding concentration of urea was 280 mg/L based upon Equations (5)–(7) (Kammonia = 1.79 × 10−5, 25 °C); thus, 280 mg/L was selected as the typical urea concentration with a decomposition temperature of 320 °C based upon the plant operating conditions. After the temperature reached the set value, the vapor and liquid phase samples in the autoclave were sampled once per hour for 8 h (the decomposition reached more than 99%). According to the results of typical urea concentration tests and the range of pH (10.0–11.0), five urea concentrations, which were 70, 140, 280, 560 and 840 mg/L, were set, and the corresponding pH was 10.29, 10.45, 10.60, 10.76 and 10.85, respectively. In the above tests, nitrogen gas was introduced into the autoclave for removing oxygen and carbon dioxide, controlling the dissolved oxygen concentration was not more than 10 μg/L. After 8 h of decomposition, the vapor and liquid phase samples were taken out from the autoclave. All samples were sealed and stored in a refrigerator at 4 °C, tested with TOC analyzer, and subjected to IR spectroscopy.

High-temperature decomposition of urea experimental device diagram.
2.3 Study of decomposition samples
The Fourier transform infrared (FT-IR) spectra of urea solution before and after decomposition were recorded using a Nicolet Avatar 360 FT-IR spectrometer. The UPLC-ESI-MS analyses of samples were performed on a Xevo Triple Quadrupole (TQ) system (Agilent, USA).
2.4 Electrochemical measurements
A conventional three-electrode system was used to evaluate the electrochemical properties of 15CrMo steel in urea and urea decomposition residual solutions with CHI660D electrochemical workstation (Chenhua Instrument Co., Ltd. Shanghai, China). It consisted of 15CrMo steel working electrode (WE) with an exposed area of 1 cm2, a platinum counter electrode and a saturated calomel reference electrode. Before the electrochemical tests, the WE was polished with 1200 grit silicon carbide papers, rinsed with deionised water, degreased in alcohol and dried under vacuum. Samples were exposed to the experimental solutions for 45 min, and the open-circuit potentials were monitored for 10 min. No substantial variation (within 10 mV per minute) of the open circuit potential confirmed the attainment of steady state.
Potentiodynamic polarization scans started at −250 mV relative to open circuit and ended at +250 mV relative to open circuit were obtained by sweeping the applied potential on the WE at constant sweep rate of 1 mV/s (Rochdi et al. 2015; Zhang et al. 2011). The corrosion current density was determined using a linear polarization method, and the corrosion rate was calculated by the following equation (Song et al. 2014):
where Icorr is the corrosion current density (A/cm2); K is combination of several conversion terms (3.2680 × 103, for mm/y); ε is the equivalent weight (27.92 g/equivalent) and ρ is the steel density (7.85 g/cm3).
Electrochemical impedance spectroscopic measurements were performed in the frequency range 100 kHz–10 mHz with AC amplitude of ±10 mV (RMS, root mean square) at the rest potential (Gama-Ferrer et al. 2012; Fernández-Domene et al. 2014). The results of impedance measurements were fitted by ZSimpWin™ software.
The electrochemical experiment was divided into two parts, where the first part is the electrochemical experiment of 15CrMo in the urea solution with different concentrations, and the second part is in the urea decomposition residue with different concentrations at 320 °C. The decomposition residue was taken for the electrochemical experiment. All the above tests were performed under unstirred conditions at room temperature of around 25 °C.
2.5 Weight loss measurements and surface analysis
The operating conditions of the power plant were simulated, and corrosion experiments were conducted in the autoclave. Weight loss measurement was used to assess the corrosion rate as a function of urea concentration. The surfaces of 15CrMo were first ground with 1200 grit silicon carbide papers, rinsed with deionized water, degreased in alcohol and dried under vacuum (Dutta et al. 2017). The concentrations of the urea solutions were 70, 140, 280, 560 and 840 mg/L. After acquiring their initial weights, the metal test pieces were suspended and soaked in 600 mL of urea solution in the autoclave, deoxygenated with nitrogen and heated at 320 °C for 8 h. After cooling to room temperature, metallic coupons were removed, scrubbed with a stiff bristle brush under running water to dislodge loosely adherent corrosion products, and soaked in cleaning solution to remove the remaining corrosion product. The preparation of the cleaning solution was as follows: 500 mL of hydrochloric acid (ρ = 1.19 g/mL) and 3.5 g of hexamethylenetetramine were mixed in distilled water to achieve a 1000 mL solution. The test pieces were cleaned repeatedly at room temperature. After the complete removal of the corrosion products, the samples were cleaned using water and acetone, dried under vacuum and reweighed [flowing Corrosion of metals and alloys–Removal of corrosion products from corrosion test specimens (Standards China 2015)]. Corrosion rate (mm/y), v, was calculated using the following relation (Wu et al. 2018):
where m0 and m1 are the weights of the metal coupons before and after weight loss tests in the urea solutions (g); S is the surface area of each metal specimen in cm2; D is the steel density (g/cm3) and T is the exposure time of the specimens in the autoclave (h).
Morphological analysis was performed to obtain an insight into the changes in the surface of the corroded samples caused by the increase in the initial concentration of urea decomposition. The elemental compositions of the samples were determined through scanning electron microscopy–energy-dispersive X-ray spectroscopy (SEM-EDS, FEI QUANTA 200). X-ray diffraction (XRD, D/max-2200/PC) measurements were performed to identify the mineralogical phases. The range of 2θ values was 10°–80° with a 0.05° step size. The scanning speed was 0.03°/s. The XRD patterns were identified using Jade + TM (Version 6.5), and a semiquantitative analysis of the compound was performed using reference intensity.
3 Results and discussion
3.1 Determination of the decomposition samples of urea solution
Figure 3 illustrates the total organic carbon (TOC) contents in the vapor and liquid phases as a function of initial urea concentration after decomposition of urea at 320 °C for 8 h. The figure shows that the TOC contents in the vapor and liquid phases after a fixed decomposition time of 8 h increase with an increase in the initial concentration of urea. The TOC content in the vapor phase is considerably larger than the TOC content in the liquid phase, indicating that the partition coefficient of TOC in the vapor phase during urea decomposition is greater than that in the liquid phase.

Contents of TOC in the vapor/liquid phase after decomposition of 280 mg/L urea at 320 °C.
Figure 4 illustrates the TOC contents of the vapor and liquid phase samples for an initial urea concentration of 280 mg/L as a function of decomposition time at 320 °C. The figure shows that the TOC in the vapor and liquid phases at typical decomposition concentration decreases with decomposition time, suggesting that the organic carbon contained in urea is converted into inorganic carbon (CO2) during the decomposition of urea. At the same time, the TOC contents in the vapor and liquid phases fluctuate in a downward trend because the NH3 and CO2 produced by urea decomposition can be converted into organic derivatives, such as ammonium carbamate, under high temperature and high pressure. As shown in Figure 4, the TOC content in the vapor phase is higher than that of the liquid phase although the TOC contents in the vapor phase and liquid phase decrease during the hydrolysis of urea, revealing that the partition coefficient of TOC in the vapor phase is greater than that in the liquid phase through the course of decomposition.

Contents of TOC in the vapor/liquid phase after decomposition of urea at 320 °C for 8 h.
Figure 5 shows the infrared spectrum of the 280 mg/L urea solution before and after decomposition at 320 °C for 8 h. The figure shows that the samples before and after decomposition have an absorption peak of –NH2 group near the wavenumber of 3440 cm−1, and N–H stretching vibration absorption peaks are observed near the wavenumber of 1600 cm−1. C–N stretching vibration absorption signals are found near 1124 and 1452 cm−1. The sample after urea decomposition exhibits a COO− absorbing vibration peak at the wavenumber of 620 cm−1 (Bacsik et al. 2011; Bacsik and Hedin 2016; Frasco 1964). The ultraperformance liquid chromatography-electrospray ionization tandem mass spectrometry (UPLC-ESI-MS) spectrum of 280 mg/L urea solution after decomposition at 320 °C for 8 h is shown in Figure 6. As shown in Figure 6, a significant peak appears at the mass-to-charge ratio of 79. Thus, the sample contains CH6N2O2. Combined with the results of Fourier transform infrared spectroscopy and UPLC-ESI-MS, the sample after urea decomposition contains ammonium carbamate (H2NCOONH4) (Equation (10)).

Infrared spectrum of 280 mg/L urea solution before and after decomposition at 320 °C for 8 h.

UPLC-ESI-MS spectrum of 280 mg/L urea solution after decomposition at 320 °C for 8 h.
The following intermediate reaction occurs during the decomposition of urea to produce NH2COO− (Equation (11)) (Prabhu-Gaunkar and Raman 1998; Qin et al. 2011; Watanabe et al. 2009):
3.2 Electrochemical characteristics of 15CrMo in urea decomposition residues
The polarization curve of 15CrMo in the urea decomposition residue is shown in Figure 7 and summarized in Table 2. Figure 7 shows that the self-corrosion potential (Ecorr) of the polarization curve negatively and gradually shifts, and icorr increases with the increase in the initial concentration of urea at ambient temperature. The corrosion rate of 15CrMo gradually increases simultaneously with the increase in corrosive substances in the decomposition residue. Figure 8 and Table 3 present the polarization curves and corresponding parameters of 15CrMo in undecomposed urea solution, respectively. The graph shows that the self-corrosion potential of 15CrMo positively and gradually shifts, and the corrosion current density decreases with the increase in urea concentration when urea solutions are not decomposed. As the same time, the corrosion rate also decreases with increasing urea concentration. It is obvious that undecomposed urea has a certain effect on corrosion inhibition. This finding is consistent with the reports in related literature studies (Çömlekçi and Ulutan 2018; Hsu et al. 2018; Ramya et al. 2016), and the corrosion of 15CrMo is caused by the decomposition products of urea.

Potentiodynamic polarization curves for 15CrMo steel in various concentrations of decomposition residue of urea solution at 320 °C.
Data of potentiodynamic polarization for 15CrMo steel decomposition residue of urea solution at 320 °C.
Conc. (mg/L) | −Ecorr (mV/SCE) | i corr (μA cm−2) | Corrosion rate (mm/y) |
---|---|---|---|
70 | 153 | 0.3917 | 3.53E-3 |
140 | 260 | 0.4944 | 5.74E-3 |
280 | 267 | 0.6499 | 7.54E-3 |
560 | 360 | 1.1403 | 1.32E-2 |
840 | 423 | 1.7428 | 2.02E-2 |

Potentiodynamic polarization curves for 15CrMo steel in various concentrations of ambient urea solution.
Data of potentiodynamic polarization curves for 15CrMo steel in various concentrations of ambient urea solution.
Conc. (mg/L) | −Ecorr (mV/SCE) | i corr (μA cm−2) | Corrosion rate (mm/y) |
---|---|---|---|
70 | 332 | 0.2887 | 3.35E-3 |
140 | 302 | 0.2693 | 3.12E-3 |
280 | 291 | 0.2424 | 2.82E-3 |
560 | 280 | 0.2126 | 2.47E-3 |
840 | 275 | 0.2118 | 2.46E-3 |
The Nyquist diagrams of 15CrMo steel in the urea decomposition residues are shown in Figure 9. And Figure 9 (a) and (b) represents Nyquist curves and Bode plots, respectively. According to the characteristics of the impedance spectrum, the equivalent circuit shown in Figure 10 was selected (Roy et al. 2014a,b), where in Rs represents the solution resistance; Rc and Cc represent the resistance and capacitance of the corrosion product film, respectively; Rct is the charge transfer resistance; CPE is a nonideal capacitance which represents the interface electric double-layer capacitor; A CPE’s impedance is given by:
where Q is the pre-exponential factor of the CPE; ω is the angular frequency; and n is the exponent, a measure of the surface irregularity of the electrode. When n = 1, 0, −1, the CPE corresponds to the capacitance (C), resistance (R) and inductance (L), respectively.

Electrochemical impedance curves for 15CrMo steel decomposition residue of urea solution exemplified as: (a) Nyquist, (b) Bode plot at 320 °C.

Equivalent circuit model used to fit the impedance spectra of 15CrMo steel in 320 °C decomposition residue of urea solution.
The parameters of each electrochemical element are listed in Table 4. The circuit has good fitness through Zsimpwin™ software fitting.
Impedance parameters of 15CrMo steel in decomposition residue of urea solution at 320 °C.
Conc. (mg/L) | R s (Ω cm−2) | C c (μF cm−2) | R ct (Ω cm−2) | Q (μS secˆn·cm−2) | n | R c (Ω cm−2) | Chi-squared |
---|---|---|---|---|---|---|---|
70 | 237.7 | 7.333E-9 | 2.409E5 | 5.351E-5 | 0.7743 | 183 | 1.741E-3 |
140 | 299.8 | 5.154E-6 | 1.206E5 | 2.933E-5 | 0.6286 | 110.2 | 5.727E-3 |
280 | 306.7 | 2.313E-9 | 9521 | 1.618E-4 | 0.5875 | 668.8 | 2.989E-3 |
560 | 146.9 | 8.516E-10 | 5117 | 1.66E-4 | 0.6798 | 2056 | 1.418E-4 |
840 | 96.3 | 3.004E-9 | 3496 | 1.861E-4 | 0.7429 | 800.4 | 5.119E-4 |
Figure 9 shows that the shapes of the impedance spectra obtained by the 15CrMo steel in the residual solutions after the decomposition tests of urea with different concentrations remain basically the same, indicating that the corrosion mechanism remains constant with the increase in urea concentration (Mahalik et al. 2010). The arcs on the real axis in the Nyquist plots are incomplete semicircles because of the roughness and nonuniformity of the 15CrMo steel surface. The impedance spectra consist of a high-frequency region capacitive reactance arc reflecting the corrosion product film information of the surface of the 15CrMo steel electrode and another large capacitive reactance arc reflecting the low-frequency region of the metal/solution interface electrochemical reaction information. The size of the semicircular diameter of the capacitive reactance arc reflects the magnitude of charge transfer resistance. The smaller the diameter of the semicircle is, the smaller the charge transfer resistance will be, indicating that the dissolution of 15CrMo steel is likely to occur, and the dissolution rate of the passivation film on the surface is accelerated. The Nyquist plots and the fitting parameters show that the corresponding capacitive antiarc radius, Rct, decreases with the increase in the initial concentration of urea, indicating that the urea decomposition products accelerate the corrosion of 15CrMo steel. The greater the initial concentration of urea is, the higher the degree of metal corrosion will be. These phenomena reveal that elevated concentrations of corrosive decomposition products promote the charge transfer reaction at the metal/solution interface (Otani et al. 2017).
3.3 Weight loss measurements and analysis
Figure 11 shows a graph of the corrosion rate trend of 15CrMo steel with initial urea concentration calculated from the weight loss data. The surface images of 15CrMo samples show that the corrosion form is localised corrosion, and the calculation result of the corrosion rate is the average value. With the increase in urea concentration, the corrosion rate of 15CrMo steels increases, which is consistent with the conclusion obtained in the electrochemical measurements. The corrosion rate of 15CrMo steel in the high-temperature corrosion test reaches 0.686 mm/y when the urea decomposition concentration is 840 mg/L.

Variation of corrosion rate of 15CrMo in urea solution at 320 °C by weight loss measurement.
Figure 12 shows the metallographic diagrams of the surfaces of 15CrMo test specimens before and after the high-temperature corrosion tests. The surface of the test piece before corrosion is only the slight scratches left after the grinding of metallographic sandpaper. After soaking and corroding in the urea solutions, the corrosion products adhere to the surface of 15CrMo steel, and the corrosion of metal test pieces becomes increasingly serious with the increase in urea concentration.

Surface images of 15CrMo samples before (a) and after (b–f) corrosion in urea solution at 320 °C.
3.4 Surface microscopic analysis
Figure 13 shows the SEM images and EDS result of 15CrMo steel before immersion in urea solution, and Figures 14 and 15 show the SEM and EDS results of the surfaces of the test specimens after high-temperature corrosion of 15CrMo steel in 560 and 840 mg/L urea solutions for 8 h. Figure 16 presents the XRD analysis of 15CrMo steel after high-temperature corrosion for 8 h in 840 mg/L urea solution. The SEM images in Figures 14(a) and 15(a) show that serious corrosion occurs on the surfaces of 15CrMo steels, and adhesion deposits of corrosion products are present after 8 h of high-temperature corrosion. The EDS analysis results in Figures 14(b) and 15(b) show that the corrosion products on the surfaces of 15CrMo steel test pieces are mainly composed of three elements, namely, C, O and Fe. The mass fraction of the three elements is more than 98%, and the atomic fraction is more than 99%. Combined with the XRD pattern in Figure 16, FeCO3 is mainly formed on the surface of 15CrMo. The XRD patterns shows that the surface of the test specimen after corrosion contains iron oxides, such as FeOOH, Fe2O3 and Fe3O4.

SEM images (a) and EDS result (b) of 15CrMo steel before immersion in urea solution.

SEM images (a) and EDS result (b) of 15CrMo steel after immersion in 560 mg/L urea solution at 320 °C for 8 h.

SEM images (a) and EDS result (b) of 15CrMo steel after immersion in 840 mg/L urea solution at 320 °C for 8 h.

XRD spectra of 15CrMo after immersion in 840 mg/L urea solution at 320 °C for 8 h.
3.5 Corrosion mechanism analysis of the failure
NH2COO− can be preferentially adsorbed on the surface of the passivating film and replace oxygen in Fe2O3. Subsequently, soluble products were generated by the combination of NH2COO− with the cations in the passivating film, resulting in the formation of small pits on the exposed base metal. The metal surface inside pitting holes was in an active state, and its electric potentials were negative. The metal surface outside the pitting holes was in a passive state, and its potential was positive. Thus, microcells were established between the active surface inside and the passive surface outside. The corrosion rate of 15CrMo remarkably increased because of the high anode current density caused by the large cathode–anode area ratio in these microcells. However, the metal surface outside the pitting holes were stabilized by cathodic protection to continuously maintain a passive state. The presence of soluble products resulting from the combination of active anions in the solution and cations in the passivation film was an essential prerequisite for the formation of pitting corrosion. In the present case, urea was heated to decomposition, and hydrated oxides enriched in Fe formed during the exposure, followed by the production of Fe2O3 in the subsequent stage of hydrothermal reaction. The alkaline environment produced by urea hydrolysis provided favorable conditions for hydrothermal precipitation. Ammonium carbamate, formed as an intermediate during urea hydrolysis, reacted with the dissolved Fe to form complexes (Krogul and Litwinienko 2015), followed by decomposition and deposition on the sample surface. As the corrosion progressed, the corrosion forms changed from pitting to overall corrosion, thereby forming a corrosion profile with pitting characteristics.
Under the action of NH2COO−, the corrosion products peeled off and migrated to where local flow was blocked and formed high deposits. The electrochemical stress corrosion cracking, caused by the original defects in the water-wall tubes under these high depositions, was the direct cause of the accident that occurred in the coal-fired power plant (Bagchi et al. 2020; Rao et al. 2016).
3.6 Prevention method of corrosion
The local design defects of the denitration system and operational errors are the main reason where the denitrification urea solution counter currently invades the thermal system. On the one hand, the improvement of the system can avoid the leakage of urea solution. For example, the compressed air pressure of the spray gun is ensured to be approximately 0.38 MPa by replacing and changing the opening angle of the spray medium of the nozzle. On the other hand, the material selection of the water-cooled wall pipe is improved. In the urea industry, ultralow carbon stainless steel has a special limit on the minimum content of Cr, Ni and Mo to enable it to obtain a fully austenitic structure with good resistance. Properly increasing the content of Nb and Mo in ferritic stainless steel can significantly improve the performance of high-temperature urea corrosion resistance (Wang et al. 2020).
4 Conclusion
Electrochemical tests showed that undecomposed urea has a slight corrosion inhibition effect of 15CrMo steel at room temperature. The initial concentration of urea has no obvious effect on the corrosion inhibition.
Urea entering the actual power thermal system decomposes to produce ammonium carbamate. Ammonium carbamate is weakly corrosive to steel at room temperature. At 320 °C, ammonium carbamate can destroy the oxide film on the surface of 15CrMo steel and cause corrosion pits on the surface of 15CrMo steel, which lead to serious degradation to the steel.
The initial concentration of urea has a significant effect on the corrosion rate of 15CrMo steel at 320 °C. The corrosion rate of 15CrMo steel increases with the increase in initial urea concentration. The maximum corrosion rate reaches 0.686 mm/y at 320 °C within a period of 8 h.
Funding source: Hunan Provincial Science and Technology Plan Key Project
Award Identifier / Grant number: 2013GK2016
-
Author contributions: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.
-
Research funding: This article was supported by Hunan Provincial Science and Technology Plan Key Project, under grant 2013GK2016.
-
Conflicts of interest: The authors declare no conflicts of interest regarding this article.
References
Bacsik, Z., Ahlsten, N., Ziadi, A., Zhao, G., Garcia-Bennett, A.E., Martín-Matute, B., and Hedin, N. (2011). Mechanisms and kinetics for sorption of CO2 on bicontinuous mesoporous silica modified with n-propylamine. Langmuir 27: 11118–11128, https://doi.org/10.1021/la202033p.Search in Google Scholar PubMed PubMed Central
Bacsik, Z. and Hedin, N. (2016). Effects of carbon dioxide captured from ambient air on the infrared spectra of supported amines. Vib. Spectrosc. 87: 215–221, https://doi.org/10.1016/j.vibspec.2016.10.006.Search in Google Scholar
Bagchi, A., Gope, D.K., Chattopadhyaya, S., and Wuriti, G. (2020). A critical review on susceptibility of stress corrosion cracking in maraging steel weldments. Mater. Today 27: 2303–2307, https://doi.org/10.1016/j.matpr.2019.09.117.Search in Google Scholar
Bernhard, A.M., Peitz, D., Elsener, M., Wokaun, A., and Kröcher, O. (2012). Hydrolysis and thermolysis of urea and its decomposition byproducts biuret, cyanuric acid and melamine over anatase TiO2. Appl. Catal. B Environ. 115–116: 129–137, https://doi.org/10.1016/j.apcatb.2011.12.013.Search in Google Scholar
Çömlekçi, G.K. and Ulutan, S. (2018). Encapsulation of linseed oil and linseed oil based alkyd resin by urea formaldehyde shell for self-healing systems. Prog. Org. Coating 121: 190–200.10.1016/j.porgcoat.2018.04.027Search in Google Scholar
Dutta, A., Saha, S.K., Adhikari, U., Banerjee, P., and Sukul, D. (2017). Effect of substitution on corrosion inhibition properties of 2-(substitutedphenyl) benzimidazole derivatives on mild steel in 1 M HCl solution: a combined experimental and theoretical approach. Corrosion Sci. 123: 256–266, https://doi.org/10.1016/j.corsci.2017.04.017.Search in Google Scholar
Eichelbaum, M., Farrauto, R.J., and Castaldi, M.J. (2010a). The impact of urea on the performance of metal exchanged zeolites for the selective catalytic reduction of NOx Part I. Pyrolysis and hydrolysis of urea over zeolite catalysts. Appl. Catal. B Environ. 97: 90–97, https://doi.org/10.1016/j.apcatb.2010.03.027.Search in Google Scholar
Eichelbaum, M., Siemer, A.B., Farrauto, R.J., and Castaldi, M.J. (2010b). The impact of urea on the performance of metal-exchanged zeolites for the selective catalytic reduction of Nox. Part II. Catalytic, FTIR, and NMR studies. Appl. Catal. B Environ. 97: 98–107, https://doi.org/10.1016/j.apcatb.2010.03.028.Search in Google Scholar
Fernández-Domene, R.M., Sánchez-Tovar, R., Escrivà-Cerdán, C., Leiva-García, R., and García-Antón, J. (2014). Study of passive films formed on AISI 316L stainless steel in non-polluted and underwater-volcano-polluted seawater. Corrosion 70: 390–401, https://doi.org/10.5006/0934.Search in Google Scholar
Forzatti, P. (2001). Present status and perspectives in de-NOx SCR catalysis. Appl. Catal. A: Gen. 222: 221–236, https://doi.org/10.1016/s0926-860x(01)00832-8.Search in Google Scholar
Frasco, D.L. (1964). Infrared spectra of ammonium carbamate and deuteroammonium carbamate. J. Chem. Phys. 41: 2134–2140, https://doi.org/10.1063/1.1726217.Search in Google Scholar
Gama-Ferrer, J.L., Gonzalez-Rodriguez, J.G., Rosales, I., and Uruchurtu, J. (2012). Corrosion behavior of Al-Sn-Bi alloys in ethylene glycol-water mixtures. Corrosion 68: 421–431, https://doi.org/10.5006/0010-9312-68.5.421.Search in Google Scholar
Hsu, C.-S. and Li, Q. (2018). Surface modification of Ti64 through hydrothermal treatment in urea solutions. Mater. Lett. 216: 299–302, https://doi.org/10.1016/j.matlet.2018.01.114.Search in Google Scholar
Koebel, M., Elsener, M., and Kleemann, M. (2000). Urea-SCR: a promising technique to reduce NOx emissions from automotive diesel engines. Catal. Today 59: 335–345, https://doi.org/10.1016/s0920-5861(00)00299-6.Search in Google Scholar
Krogul, A. and Litwinienko, G. (2015). One pot synthesis of ureas and carbamates via oxidative carbonylation of aniline-type substrates by CO/O2 mixture catalyzed by Pd-complexes. J. Mol. Catal. A: Chem 407: 204–211, https://doi.org/10.1016/j.molcata.2015.06.027.Search in Google Scholar
Kröcher, O., Elsener, M., and Koebel, M. (2005). An ammonia and isocyanic acid measuring method for soot containing exhaust gases. Anal. Chim. Acta 537: 393–400, https://doi.org/10.1016/j.aca.2004.12.082.Search in Google Scholar
Lu, J., Yang, Z., Zhang, B., Huang, J., and Xu, H. (2018). Corrosion behavior of candidate materials used for urea hydrolysis equipment in coal-fired selective catalytic reduction units. J. Ind. Eng. Chem. 27: 3290–3296, https://doi.org/10.1007/s11665-017-3001-3.Search in Google Scholar
Lu, J., Zhang, B., Huang, J., Yang, Z., and Xu, H. (2017). Corrosion behavior of candidate materials used for critical parts of urea hydrolysis equipment in gas denitrification. Mater. Mech. Eng. 41: 6–12.Search in Google Scholar
Liu, J., Song, J., Xiao, H., Zhang, L., Qin, Y., Liu, D., Hou, W., and Du, N. (2014). Synthesis and thermal properties of ZnAl layered double hydroxide by urea hydrolysis. Powder Technol. 253: 41–45, https://doi.org/10.1016/j.powtec.2013.11.007.Search in Google Scholar
Mahalik, K., Sahu, J.N., Patwardhan, V., and Meikap, B.C. (2010). Statistical modelling and optimization of hydrolysis of urea to generate ammonia for flue gas conditioning. J. Hazard Mater. 182: 603–610, https://doi.org/10.1016/j.jhazmat.2010.06.075.Search in Google Scholar PubMed
Nockert, J. and Norell, M. (2014). Corrosion at the urea injection in SCR-system during component test. Mater. Corros. 64: 34–42.10.1002/maco.201005983Search in Google Scholar
Otani, K., Sakairi, M., and Islam, M.S. (2017). Influence of metal cations on inhibitor performance of gluconates in the corrosion of mild steel in fresh water. Corrosion Rev. 36: 105–113.10.1515/corrrev-2017-0047Search in Google Scholar
Prabhu-Gaunkar, G.V. and Raman, A. (1998). Corrosion inhibitors in fertilizer production and handling. Corrosion Rev. 16: 393, https://doi.org/10.1515/corrrev.1998.16.4.393.Search in Google Scholar
Qin, F., Wang, S., Kim, I., Svendsen, H.F., and Chen, C. (2011). Heat of absorption of CO2 in aqueous ammonia and ammonium carbonate/carbamate solutions. Int. J. Greenh. Gas Contr. 5: 405–412. https://doi.org/10.1016/j.ijggc.2010.04.005.Search in Google Scholar
Ramya, K., Anupama, K.K., Shainy, K.M., and Abraham, J. (2016). Synergistic and hydrogen bonded interaction of alkyl benzimadazoles and urea pair on mild steel in hydrochloric acid: adsorption, electroanalytical and theoretical studies. J. Taiwan Inst. Chem. Eng. 58: 517–527.10.1016/j.jtice.2015.05.036Search in Google Scholar
Rao, A.C.U., Vasu, V., Govindaraju, M., and Srinadh, K.V.S. (2016). Stress corrosion cracking behaviour of 7xxx aluminum alloys: a literature review. Trans. Nonferr. Metal Soc. 26: 1447–1471, https://doi.org/10.1016/s1003-6326(16)64220-6.Search in Google Scholar
Rochdi, A., Touir, R., Bakri, M.E., Touhami, M.E., Bakkali, S., and Mernari, B. (2015). Protection of low carbon steel by oxadiazole derivatives and biocide against corrosion in simulated cooling water system. J. Environ. Chem. Eng. 3: 233–242, https://doi.org/10.1016/j.jece.2014.11.020.Search in Google Scholar
Roy, P., Karfa, P., Adhikari, U., and Sukul, D. (2014a). Corrosion inhibition of mild steel in acidic medium by polyacrylamide grafted Guar gum with various grafting percentage: effect of intramolecular synergism. Corrosion Sci. 88: 246–253, https://doi.org/10.1016/j.corsci.2014.07.039.Search in Google Scholar
Roy, P., Pal, A., and Sukul, D. (2014b). Origin of the synergistic effect between polysaccharide and thiourea towards adsorption and corrosion inhibition for mild steel in sulphuric acid. RSC Adv. 4: 10607–10613, https://doi.org/10.1039/c3ra46549g.Search in Google Scholar
Sahu, J.N., Chava, V.S.R.K., Hussain, S., Patwardhan, A.V., and Meikap, B.C. (2010). Optimization of ammonia production from urea in continuous process using ASPEN Plus and computational fluid dynamics study of the reactor used for hydrolysis process. J. Ind. Eng. Chem. 16: 577–586, https://doi.org/10.1016/j.jiec.2010.03.016.Search in Google Scholar
Shaikh, H., Rao, R.V.S., George, R.P., Anita, T., and Khatak, H.S. (2003). Corrosion failures of AISI type 304 stainless steel in a fertilizer plant. Eng. Fail. Anal. 10: 329–339. https://doi.org/10.1016/s1350-6307(02)00076-6.Search in Google Scholar
Song, G.-D., Kim, M.-H., and Maeng, W.-Y. (2014). Optimization of polymeric dispersant concentration for the dispersion-stability of magnetite nanoparticles in water solution. J. Nanosci. Nanotechnol. 14: 9525–9533, https://doi.org/10.1166/jnn.2014.10174.Search in Google Scholar PubMed
Standards China. (2015). Corrosion of metals and alloys-removal of corrosion products from corrosion test specimens, (ISO:8407:2009). IS Standards Press of China, Beijing.Search in Google Scholar
Starostin, M., Shter, G.E., and Grader, G.S. (2016). Corrosion of aluminum alloys Al 6061 and Al 2024 in ammonium nitrate-urea solution. Mater. Corros. 67: 387–395, https://doi.org/10.1002/maco.201508552.Search in Google Scholar
Todorova, T., Peitz, D., Kröcher, O., Wokaun, A., and Delley, B. (2010). Guanidinium formate decomposition on the (101) TiO2-anatase surface: combined minimum energy reaction pathway calculations and temperature-programmed decomposition experiments. J. Phys. Chem. C 115: 1195–1203, https://doi.org/10.1021/jp106523b.Search in Google Scholar
Watanabe, H., Yamazaki, H., Wang, X., and Uchiyama, S. (2009). Oxygen and hydrogen peroxide reduction catalyses in neutral aqueous media using copper ion loaded glassy carbon electrode electrolyzed in ammonium carbamate solution. Electrochim. Acta 54: 1362–1367, https://doi.org/10.1016/j.electacta.2008.09.020.Search in Google Scholar
Wang, X.L., Huang, A.R., Li, M.X., Zhang, W., Shang, C.J., Wang, J.L., and Xie, Z.J. (2020). The significant roles of Nb and Mo on enhancement of high temperature urea corrosion resistance in ferritic stainless steel. Mater. Lett. 269: 127660, https://doi.org/10.1016/j.matlet.2020.127660.Search in Google Scholar
Wu, T., Wu, H., Niu, G., Li, T., Sun, R., Gu, Y., and Yuan, R. (2018). Effect of microstructure on the corrosion performance of 5% Cr steel in a CO2 environment. Corrosion 74: 757–767.10.5006/2595Search in Google Scholar
Zhang, G.A. and Cheng, Y.F. (2011). Localized corrosion of carbon steel in a CO2-saturated oilfield formation water. Electrochim. Acta 56: 1676–1685, doi:https://doi.org/10.1016/j.electacta.2010.10.059.Search in Google Scholar
Zhang, X., Zhang, B., Lu, X., Gao, N., Xiang, X., and Xu, H. (2017). Experimental study on urea hydrolysis to ammonia for gas denitration in a continuous tank reactor. Energy 126: 677–688, https://doi.org/10.1016/j.energy.2017.03.067.Search in Google Scholar
© 2020 Walter de Gruyter GmbH, Berlin/Boston
Articles in the same Issue
- Frontmatter
- Reviews
- Corrosion of rail tracks and their protection
- Electrogalvanization using new generation coatings with carbonaceous additives: progress and challenges
- Phytochemicals as steel corrosion inhibitor: an insight into mechanism
- Original articles
- Corrosion behavior of 15CrMo steel for water-wall tubes in thermal power plants in the presence of urea and its byproducts
- Stir zone stress corrosion cracking behavior of friction stir welded AA7075-T651 aluminum alloy joints
- Optimization of erosion wears of Al–Mg–Si–Cu–SiC composite produced by the PM method
- Annual reviewer acknowledgement
- Reviewer acknowledgement Corrosion Reviews volume 38 (2020)
Articles in the same Issue
- Frontmatter
- Reviews
- Corrosion of rail tracks and their protection
- Electrogalvanization using new generation coatings with carbonaceous additives: progress and challenges
- Phytochemicals as steel corrosion inhibitor: an insight into mechanism
- Original articles
- Corrosion behavior of 15CrMo steel for water-wall tubes in thermal power plants in the presence of urea and its byproducts
- Stir zone stress corrosion cracking behavior of friction stir welded AA7075-T651 aluminum alloy joints
- Optimization of erosion wears of Al–Mg–Si–Cu–SiC composite produced by the PM method
- Annual reviewer acknowledgement
- Reviewer acknowledgement Corrosion Reviews volume 38 (2020)